WO2020112843A1 - Solid state batteries - Google Patents

Solid state batteries Download PDF

Info

Publication number
WO2020112843A1
WO2020112843A1 PCT/US2019/063354 US2019063354W WO2020112843A1 WO 2020112843 A1 WO2020112843 A1 WO 2020112843A1 US 2019063354 W US2019063354 W US 2019063354W WO 2020112843 A1 WO2020112843 A1 WO 2020112843A1
Authority
WO
WIPO (PCT)
Prior art keywords
lgps
rechargeable battery
solid state
mpa
battery
Prior art date
Application number
PCT/US2019/063354
Other languages
French (fr)
Other versions
WO2020112843A8 (en
Inventor
Luhan YE
William Fitzhugh
Fun WU
Xin Li
Original Assignee
President And Fellows Of Harvard College
Priority date (The priority date is an assumption and is not a legal conclusion. Google has not performed a legal analysis and makes no representation as to the accuracy of the date listed.)
Filing date
Publication date
Application filed by President And Fellows Of Harvard College filed Critical President And Fellows Of Harvard College
Priority to CN201980090173.8A priority Critical patent/CN113454825A/en
Priority to US17/297,228 priority patent/US20210408580A1/en
Priority to CA3120864A priority patent/CA3120864A1/en
Priority to AU2019387113A priority patent/AU2019387113A1/en
Priority to JP2021529271A priority patent/JP2022509633A/en
Priority to KR1020217019956A priority patent/KR20210100651A/en
Priority to EP19890968.1A priority patent/EP3888175A4/en
Publication of WO2020112843A1 publication Critical patent/WO2020112843A1/en
Publication of WO2020112843A8 publication Critical patent/WO2020112843A8/en

Links

Classifications

    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M10/00Secondary cells; Manufacture thereof
    • H01M10/05Accumulators with non-aqueous electrolyte
    • H01M10/056Accumulators with non-aqueous electrolyte characterised by the materials used as electrolytes, e.g. mixed inorganic/organic electrolytes
    • H01M10/0561Accumulators with non-aqueous electrolyte characterised by the materials used as electrolytes, e.g. mixed inorganic/organic electrolytes the electrolyte being constituted of inorganic materials only
    • H01M10/0562Solid materials
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M10/00Secondary cells; Manufacture thereof
    • H01M10/05Accumulators with non-aqueous electrolyte
    • H01M10/058Construction or manufacture
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M10/00Secondary cells; Manufacture thereof
    • H01M10/05Accumulators with non-aqueous electrolyte
    • H01M10/052Li-accumulators
    • H01M10/0525Rocking-chair batteries, i.e. batteries with lithium insertion or intercalation in both electrodes; Lithium-ion batteries
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M10/00Secondary cells; Manufacture thereof
    • H01M10/42Methods or arrangements for servicing or maintenance of secondary cells or secondary half-cells
    • H01M10/44Methods for charging or discharging
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M4/00Electrodes
    • H01M4/02Electrodes composed of, or comprising, active material
    • H01M4/36Selection of substances as active materials, active masses, active liquids
    • H01M4/48Selection of substances as active materials, active masses, active liquids of inorganic oxides or hydroxides
    • H01M4/485Selection of substances as active materials, active masses, active liquids of inorganic oxides or hydroxides of mixed oxides or hydroxides for inserting or intercalating light metals, e.g. LiTi2O4 or LiTi2OxFy
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M4/00Electrodes
    • H01M4/02Electrodes composed of, or comprising, active material
    • H01M4/36Selection of substances as active materials, active masses, active liquids
    • H01M4/48Selection of substances as active materials, active masses, active liquids of inorganic oxides or hydroxides
    • H01M4/50Selection of substances as active materials, active masses, active liquids of inorganic oxides or hydroxides of manganese
    • H01M4/505Selection of substances as active materials, active masses, active liquids of inorganic oxides or hydroxides of manganese of mixed oxides or hydroxides containing manganese for inserting or intercalating light metals, e.g. LiMn2O4 or LiMn2OxFy
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M4/00Electrodes
    • H01M4/02Electrodes composed of, or comprising, active material
    • H01M4/36Selection of substances as active materials, active masses, active liquids
    • H01M4/48Selection of substances as active materials, active masses, active liquids of inorganic oxides or hydroxides
    • H01M4/52Selection of substances as active materials, active masses, active liquids of inorganic oxides or hydroxides of nickel, cobalt or iron
    • H01M4/525Selection of substances as active materials, active masses, active liquids of inorganic oxides or hydroxides of nickel, cobalt or iron of mixed oxides or hydroxides containing iron, cobalt or nickel for inserting or intercalating light metals, e.g. LiNiO2, LiCoO2 or LiCoOxFy
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M4/00Electrodes
    • H01M4/02Electrodes composed of, or comprising, active material
    • H01M4/36Selection of substances as active materials, active masses, active liquids
    • H01M4/58Selection of substances as active materials, active masses, active liquids of inorganic compounds other than oxides or hydroxides, e.g. sulfides, selenides, tellurides, halogenides or LiCoFy; of polyanionic structures, e.g. phosphates, silicates or borates
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M2300/00Electrolytes
    • H01M2300/0017Non-aqueous electrolytes
    • H01M2300/0065Solid electrolytes
    • H01M2300/0068Solid electrolytes inorganic
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M2300/00Electrolytes
    • H01M2300/0017Non-aqueous electrolytes
    • H01M2300/0065Solid electrolytes
    • H01M2300/0068Solid electrolytes inorganic
    • H01M2300/008Halides
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01MPROCESSES OR MEANS, e.g. BATTERIES, FOR THE DIRECT CONVERSION OF CHEMICAL ENERGY INTO ELECTRICAL ENERGY
    • H01M4/00Electrodes
    • H01M4/02Electrodes composed of, or comprising, active material
    • H01M4/36Selection of substances as active materials, active masses, active liquids
    • H01M4/362Composites
    • H01M4/366Composites as layered products
    • YGENERAL TAGGING OF NEW TECHNOLOGICAL DEVELOPMENTS; GENERAL TAGGING OF CROSS-SECTIONAL TECHNOLOGIES SPANNING OVER SEVERAL SECTIONS OF THE IPC; TECHNICAL SUBJECTS COVERED BY FORMER USPC CROSS-REFERENCE ART COLLECTIONS [XRACs] AND DIGESTS
    • Y02TECHNOLOGIES OR APPLICATIONS FOR MITIGATION OR ADAPTATION AGAINST CLIMATE CHANGE
    • Y02EREDUCTION OF GREENHOUSE GAS [GHG] EMISSIONS, RELATED TO ENERGY GENERATION, TRANSMISSION OR DISTRIBUTION
    • Y02E60/00Enabling technologies; Technologies with a potential or indirect contribution to GHG emissions mitigation
    • Y02E60/10Energy storage using batteries
    • YGENERAL TAGGING OF NEW TECHNOLOGICAL DEVELOPMENTS; GENERAL TAGGING OF CROSS-SECTIONAL TECHNOLOGIES SPANNING OVER SEVERAL SECTIONS OF THE IPC; TECHNICAL SUBJECTS COVERED BY FORMER USPC CROSS-REFERENCE ART COLLECTIONS [XRACs] AND DIGESTS
    • Y02TECHNOLOGIES OR APPLICATIONS FOR MITIGATION OR ADAPTATION AGAINST CLIMATE CHANGE
    • Y02PCLIMATE CHANGE MITIGATION TECHNOLOGIES IN THE PRODUCTION OR PROCESSING OF GOODS
    • Y02P70/00Climate change mitigation technologies in the production process for final industrial or consumer products
    • Y02P70/50Manufacturing or production processes characterised by the final manufactured product

Definitions

  • the invention is directed to the field of solid state batteries with alkali metal sulfide solid state electrolytes.
  • Solid-state lithium ion conductors the key component to enabling all solid-state lithium ion batteries, are one of the most pursued research objectives in the battery field.
  • the intense interest in solid-state electrolytes, and solid-state batteries more generally, stems principally from improved safety, the ability to enable new electrode materials and better low-temperature performance.
  • Safety improvements are expected for solid-state battery cells as the currently used liquid-electrolytes are typically highly-flammable organic solvents. Replacing these electrolytes with non-flammable solids would eliminate the most problematic aspect of battery safety.
  • solid-electrolytes are compatible with several high energy density electrode materials that cannot be implemented with liquid-electrolyte based configurations.
  • Solid-electrolytes also maintain better low temperature operation than liquid-electrolytes, which experience substantial ionic conductivity drops at low temperatures. Such low temperature performance is critical in the burgeoning electric-vehicles market. Of the currently studied solid-electrolytes, sulfides remain one of the highest-performance and most promising families. Sulfide glass solid-electrolytes and glass-ceramic solid-electrolytes, where crystalline phases have precipitated within a glassy matrix, have demonstrated ionic conductivities on the order of 0 ?
  • LGPS was one of the first solid-electrolytes to reach ionic conductivities comparable to liquid-electrolytes at only to be displaced by LSPS, which achieved an astonishingly high ionic conductivity of Despite these promising
  • rechargeable solid state batteries using solid state electrolytes with improved cycling performance.
  • the rechargeable solid state batteries disclosed herein are advantageous as the solid state electrolytes have superior voltage stability and excellent battery cycle performance. Batteries of the invention may be stabilized against the formation of lithium dendrites and/or can operate at high current density for an extended number of cycles.
  • the invention features a rechargeable battery including a first electrode, a second electrode, and a solid state electrolyte disposed therebetween.
  • the solid state electrolyte includes a sulfide that includes an alkali metal, such as lithium.
  • the solid state electrolyte is under a volumetric constraint sufficient to stabilize the solid state electrolyte during electrochemical cycling.
  • the volumetric constraint exerts a pressure of about 70 to about 1,000 MPa, e.g., about 100-250 MPa, on the solid state electrolyte, e.g., to enforce mechanical constriction on the microstructure of solid electrolyte on the order of 15 GPa.
  • the volumetric constraint provides a voltage stability window of between 1 and 10 V, e.g., 1-8V, 5.0-8 V, or greater than 5.7 V, or even greater than 10V.
  • the solid state electrolyte has a core shell morphology.
  • the alkali metal is Li, Na, K, Rb, or Cs, e.g., Li.
  • the solid state electrolyte includes SiPS, GePS, SnPS, PSI, or PS.
  • the solid state electrolyte is Li10SiP2S12, Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3.
  • the first electrode is the cathode, which can include LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or LiCo0.5Mn1.5O4.
  • the second electrode is anode and can include lithium metal, lithiated graphite, or Li4Ti5O12.
  • the volumetric constraint provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa, e.g., about 15 GPa.
  • the invention features a rechargeable battery including a first electrode, a second electrode, and a solid state electrolyte disposed therebetween, wherein the second electrode is an anode comprising an alkali metal and graphite.
  • the battery is under a pressure of about 70-1000 MPa, e.g., about 100-250 MPa.
  • the alkali metal and graphite form a composite.
  • the alkali metal is Li, Na, K, Rb, or Cs, e.g., Li.
  • the solid state electrolyte includes SiPS, GePS, SnPS, PSI, or PS.
  • the solid state electrolyte is Li10SiP2S12, Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3.
  • the first electrode is the cathode and can include LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or LiCo0.5Mn1.5O4.
  • the battery is under an external stress that provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa, e.g., about 15 GPa.
  • the invention features a rechargeable battery including a first electrode, a second electrode, and a solid state electrolyte disposed therebetween, wherein the solid state electrolyte may include a sulfide including an alkali metal; and the battery is under isovolumetric constraint.
  • the isovolumetric constraint is provided by compressing the solid state electrolyte under a pressure of about 3-1000 MPa, e.g., about 100-250 MPa.
  • the alkali metal is Li, Na, K, Rb, or Cs, e.g., Li.
  • the solid state electrolyte includes SiPS, GePS, SnPS, PSI, or PS.
  • the solid state electrolyte is Li10SiP2S12, Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3.
  • the first electrode is the cathode and can include LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or LiCo0.5Mn1.5O4.
  • the isovolumetric constraint provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa, e.g., about 15 GPa.
  • the invention features a rechargeable battery having a first electrode, a second electrode, and a solid state electrolyte disposed therebetween.
  • the solid state electrolyte includes a sulfide that includes an alkali metal, and optionally has a core-shell morphology.
  • the first electrode includes an interfacially stabilizing coating material.
  • the first and second electrodes independently include an interfacially stabilizing coating material.
  • one of the first and second electrodes includes a lithium-graphite composite.
  • the first electrode comprises a material as described herein, e.g., in Table 1.
  • the coating material of the first electrode is a coating material as described herein, e.g., LiNbO3, AlF3, MgF2, Al2O3, SiO2, graphite, or in Table 2.
  • the alkali metal is Li, Na, K, Rb, or Cs, e.g., Li.
  • the solid state electrolyte includes SiPS, GePS, SnPS, PSI, or PS.
  • the solid state electrolyte is Li10SiP2S12, Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3.
  • the first electrode is the cathode and can include LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or
  • the battery is under an external stress that provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa, e.g., about 15 GPa. In certain embodiments, the battery is under a pressure of about 70-1000 MPa, e.g., about 100-250 MPa.
  • the invention features a method of storing energy by applying a voltage across the first and second electrodes and charging the rechargeable battery of the invention.
  • the invention provides a method of providing energy by connecting a load to the first and second electrodes and allowing the rechargeable battery of the invention to discharge.
  • FIGS 1A-1B Cyclic Voltammetry (CV) tests of LGPS in liquid (A) and solid (B) states at different pressures.
  • LGPS/C thin film with the ratio of 90:10 was tested in the liquid electrolyte (black curve in (A)).
  • the CV tests were also conducted by replacing liquid electrolyte with LGPS pellets, which is all- solid-state CV, at different pressures.
  • the decomposition intensity is decreased significantly with increasing applied pressure.
  • At a reasonably low pressure of 6 T (420 MPa) there is already no notable decomposition peaks before 5.7 V (purple curve), which indicates applying external pressure or volume constriction on the battery cell level can widen the electrochemical window of the solid- state electrolyte.
  • FIGS 2A-2B Capacity (A) and cycling performance (B) of LiCoO2 (LCO)- Li4Ti5O12 (LTO) all-solid- state full battery.
  • LCO LiCoO2
  • LTO Li4Ti5O12
  • Figures 3A-3B Capacity (A) and cycling performance (B) of LiNi0.5Mn1.5O4 (LNMO)-LTO all-solid- state full battery.
  • LTO LiNi0.5Mn1.5O4
  • B cycling performance
  • Figure 4 High voltage cathode candidates for 6V and greater all solid state Li-ion battery technology.
  • the legend labels are: F are fluorides, O are oxides, P,O are phosphates, and S,O: sulfates.
  • the complete list of these high voltage fluorides, oxides, phosphates, and sulfates is provided in Table 1.
  • Commercial LiCoO2 (LCO) and LMNO are labeled as stars.
  • Figs 5A-5B (A) Illustration of the impact of strain on LGPS decomposition, where ? ? is the fraction of LGPS that has decomposed.
  • the lower dashed line represents the Gibbs energy (? ? ?? ? ?) of a binary combination of pristine LGPS and an arbitrary set of decay products (?) when negligible pressure is applied (isobaric decay with The solid line shows the Gibbs when a mechanical constraint is applied to the LGPS. Since LGPS tends to expand upon decomposition, the strain Gibbs (? ?????? ) increases when such a mechanical constraint is applied. At some fracture point, denoted ? ? , the Gibbs energy of the system exceeds the energy needed to fracture the mechanical constraints (the upper dashed line). The highlighted path is the suggested ground state for a mechanically constrained LGPS system. The region is metastable if Schematic representation of work
  • FIGS 8A-8E Voltage (j), lithium chemical potential ( ) and Fermi level distributions in various battery configurations.
  • A Conventional battery design.
  • B Conventional battery with hybrid solid-electrolyte/active material cathode.
  • c ? gives the interface voltage that forms between the active material and the solid-electrolyte because of the different lithium ion chemical potentials.
  • C Illustration of previous speculation of how insulating layers could lead to variable lithium metal chemical potentials within the cell.
  • D Expectation of how the voltage from part (C) would relax given the effective electronic conduction that occurs due to lithium hole migration.
  • E The result of part (D) once the applied voltage exceeds the intrinsic stability window of the solid-electrolyte.
  • FIGS 9A-9D Comparison between microstructures and chemical composition of LGPS and ultra- LGPS particles.
  • A, C Typical TEM bright-field images of LGPS and ultra-LGPS particles respectively, showing a distinct surface layer for ultra-LGPS particle.
  • B, D Statistically analyzed STEM EDS linescans performed on various LGPS and ultra-LGPS particles with different sizes, showing a uniform distribution of sulfur concentration from surface to bulk for LGPS particles, but a decreased sulfur concentration in surface layer for ultra-LGPS.
  • Figure 10 STEM EDS linescans across individual LGPS particles with different particle sizes ranging from 100nm to 3 ⁇ m, showing that the sulfur concentration variation from surface to the bulk has no regular pattern.
  • FIG. 11 STEM EDS linescans across individual LGPS particles sonicated in dimethyl carbonate (DMC) for 70h with different particle sizes ranging from 60nm to 4 ⁇ m, showing that sulfur concentration is obviously smaller at surface region compared to that in the bulk.
  • DMC dimethyl carbonate
  • Figure 12 STEM EDS linescans across individual LGPS particles sonicated in diethyl carbonate (DEC) for 70h with different particle sizes ranging from 120nm to 4 ⁇ m, showing that sulfur concentration is obviously smaller at surface region compared to that in the bulk.
  • DEC diethyl carbonate
  • FIG. 13 Quantitative STEM EDX analyses of LGPS particles before and after ultrasonic preparation show that surface/bulk ratio of S is obviously lower after sonication in organic electrolytes (DEC and DMC).
  • Figure 14 STEM EDS linescans across individual LGPS particles soaked in DMC for 70h without sonication with different particle sizes ranging from 160nm to 3 ⁇ m, showing that the sulfur concentration variation from surface to the bulk has no regular pattern.
  • Figures 15A-15H Comparison between electrochemical performances of LGPS and ultra-LGPS particles, and LIBs made from LGPS and ultra-LGPS particles.
  • A, B Cyclic voltammograms(CV) of Li/LGPS/LGPS+C/Ta and Li/ultra-LGPS/ultra-LGPS/Ta cells respectively, with a lithium reference electrode at a scan rate of 0.1mVs -1 and a scan range of 0.5 to 5 V.
  • C, D Sensitive electrochemical impedance spectra (EIS) for LGPS and ultra-LGPS cells in panel (A,B) before and after CV tests.
  • E, F Charge-discharge profiles of LGPS-LIB (LTO+LGPS+C/Glass fiber separator/Li) and ultra-LGPS- LIB (LTO+ultra-LGPS+C/Glass fiber separator/Li) cycled at 0.5C current rate in the voltage range of 1.0 - 2.2 V.
  • G, H Cyclic capacity curves of LGPS LIB and ultra-LGPS-LIB.
  • FIGs 16A-16B Cycling performance of (A) LGPS-ASSLIB (LTO+LGPS+C as cathode, LGPS as solid electrolyte, and Li as anode) and (B) ultra-LGPS-ASSLIB (LTO+ultra-LGPS+C as cathode, ultra- LGPS as solid electrolyte, and Li as anode) at low current rate (0.02C).
  • Figures 17A-17B Cycling performance of (A) LGPS-ASSLIB (LTO+LGPS+C as cathode, LGPS as solid electrolyte, and Li as anode) and (B) ultra-LGPS-ASSLIB (LTO+ultra-LGPS+C as cathode, ultra- LGPS as solid electrolyte, and Li as anode) at medium current rate (0.1C).
  • Figure 18A-18B Cycling performance of (A) LGPS-ASSLIB (LTO+LGPS+C as cathode, LGPS as solid electrolyte, and Li as anode) and (B) ultra-LGPS-ASSLIB (LTO+ultra-LGPS+C as cathode, ultra- LGPS as solid electrolyte, and Li as anode) at high current rate (0.8C).
  • FIGS 19A-19G Microstructural and compositional (S)TEM studies of LTO/LGPS interfaces after cycling in LGPS ASSLIB.
  • A FIB sample prepared from LGPS ASSLIB after 1 charge- discharge cycle, in which the cathode layer (LTO+LGPS+C) and SE layer (LGPS) are included.
  • B TEM BF images of LTO/LGPS primary interface, showing a transit layer with multiple dark particles.
  • C HRTEM image of LTO particle and its corresponding FFT pattern.
  • STEM DF image of LTO/LGPS primary interface shows super bright particles within the transit layer, indicating the accumulation of heavy elements.
  • STEM EELS linescans performed across the primary interface, indicating that the bright particles within the transit layer are sulfur-rich.
  • F STEM DF image of LTO/LGPS secondary interface, in which a higher density of bright particles with similar morphology show up again.
  • G STEM EELS linescans performed across the secondary interface, indicating that the bright particles are sulfur-rich.
  • Figure 20 TEM bright-field images and STEM dark-field image of primary LTO/LGPS interface (interface between cathode and LGPS solid electrolyte layer) of LGPS-ASSLIB (LTO+LGPS+C as cathode, LGPS as solid electrolyte, and Li as anode), showing an obvious transit layer between the cathode and solid electrolyte layer.
  • Figures 21A-21B (A) STEM dark-field image of and (B) EELS linescan on primary LTO/LGPS interface (interface between cathode and LGPS solid electrolyte layer) of LGPS-ASSLIB
  • Figures 22A-22B (A) STEM dark-field image of and (B) EELS linescan on primary LTO/LGPS interface (interface between cathode and LGPS solid electrolyte layer) of LGPS-ASSLIB
  • Figures 23A-23F Microstructural and compositional (S)TEM studies of LTO/ultra-LGPS interfaces after cycling in ultra-LGPS ASSLIB.
  • A TEM BF image of LTO/ultra-LGPS primary interface, showing a smooth interface with no dark particles that exist in Figure 6B.
  • B STEM EELS linescan spectra corresponding to the dashed arrow in Figure 23A.
  • C STEM DF image of LTO/ultra-LGPS secondary interface.
  • D STEM EDS linescans show a continuously decreasing atomic percentage of sulfur from inner ultra-LGPS particle to secondary LTO/ultra-LGPS interface, and finally into LTO+C composite region.
  • E STEM EDS mapping shows that the large particle in Figure 22C is LGPS particle.
  • Figure 25A-25C (A) The number of hulls required to evaluate the stability of the 67k materials considered if the evaluation schema is material iteration (left columns) or elemental set iteration (right columns). (B) An illustration of the pseudo-binary approach to interfacial stability between LSPS and an arbitrary material A. represents the materials-level decomposition energy that exists even in
  • Figures 26A-26C (A-C) Correlation of elemental species fraction with the added electrochemical interfacial instability ( ???? ) at 0, 2 and 4 V, respectively. Negative values are those species such that increasing concentration decreases and improves interfacial stability. Conversely, positive values are those species that tend to increase and worsen interfacial stability. Elements that are
  • FIGS 27A-27D (A) Hull energy vs voltage relative to lithium metal for LSPS. Darker Gray [Mid- Gray] shading highlights where the decomposition is oxidative [reductive]. Light gray shading represents the region where LSPS decays to without consuming or producing lithium (e.g. lithium neutral). The oxidation [reduction] region is characterized by a hull energy that increases [decreases] with increasing voltage. (B) and (C) Hull energies at the boundary voltages for the anode and cathode ranges, respectively, in terms of anionic species (e.g., oxygen containing compounds vs sulfur containing compounds, etc.). Data points below [above] the neutral decay line are net oxidative
  • Figures 28A-28C Comparison of average LSPS interfacial stability of compounds sorted by anionic species. (A) The average total maximum kinetic driving energy and the contribution due to
  • Figures 29A-29B Functionally stable results for compounds sorted by anionic species.
  • A) and B The total number (line) and percentage (bar) of each anionic class that was determined to be functionally stable.
  • the bottom bar represented the percentage of materials that are functionally stable and the top bar represents the percentage of materials that are potentially functionally stable depending on the reversibility of lithiation/delithiation.
  • Figures 30A-30F Comparison of XRD patterns to show structural decay of LCO, SnO2, LTO and SiO2 at the solid-electrolyte material interface (with no applied voltage).
  • ⁇ , , , , ⁇ stand for LCO(PDF# 44-0145), LSPS(ICSD#252037), SiO2(PDF# 48-0476), Li3PO4(PDF# 45-0747), Cubic Co4S3(PDF# 02-1338), Monoclinic Co4S3(PDF# 02-1458) respectively.
  • Figures 31A-31E Comparison of XRD patterns for each individual phase: (A) LiCoO2, (B) LSPS, (C) Li4Ti5O12, (D) SnO2 and (E) SiO2, at room temperature and 500°C. No significant change between room temperature and 500°C can be observed for each phase.
  • Figures 32A-32D Comparison of XRD patterns for mixture powders: (A) LiCoO2+LSPS, (B)
  • the onset reaction temperature is observed to be 500°C, 400°C and 500°C for LiCoO2+LSPS, SnO2+LSPS and Li4Ti5O12+LSPS, respectively. No reaction is observed to happen for SiO2+LSPS up to 500°C.
  • Figures.33A-33F (A, B, C) XRD of different powder mixtures before and after heat treatment at 500°C for 36 hours ((A) Li + LGPS; (B) Graphite + LGPS; (C) Lithiated graphite + LGPS).
  • the symbols and corresponding phases are: ) The structure of Li/Graphite anode in LGPS based all-solid-state battery; (E) SEM image of the cross section of Li/Graphite anode; (F) FIB-SEM of the interface of Li and Graphite.
  • FIGS.34A-34E (A) The comparison of cyclic performance between Li/G-LGPS-G/Li and Li-LGPS-Li symmetric batteries; (B) The SEM images of symmetric batteries after cycling. Li/G-LGPS-G/Li symmetric battery after 300 hours’ cycling (B1,2) and Li-LGPS-Li symmetric battery after 10 hours’ cycling (B3,4); (C) The rate performance of Li/G-LGPS-G/Li symmetric batteries under different pressures. (D) The SEM images of Li/G-LGPS-G/Li symmetric batteries under different pressures after rate tests. (E) The ultra-high rate performance up to 10 mA/cm 2 of Li/G-LGPS-G/Li symmetric batteries.
  • the pressure applied in (E) is 250 MPa.
  • Insets are the cycling profiles plotted in the range of -0.3V to 0.3V, showing that there is no obvious change of overpotential after high rate cycling. More voltage profile enlargements are shown in supplementary information Figure 42.
  • Figures.35A-35D (A) The comparison of initial charge/ discharge curves, (B) the initial Coulombic efficiencies and (C) the open circuit voltages after 1h rest, among different capacity ratios of Li to Graphite in Li/G-LGPS-LCO (LiNbO3 coated) system.
  • the Li/G capacity ratio of 0, 0.5, 0.8, 1.5, 2.5 and 4 can be translated into Li/G thickness ratio of around 0, 0.3, 0.4, 0.8, 1.3, and 2.1 respectively. Without specific explanation, the Li/graphite thickness ratio is 1.0-1.3 by default in this work.
  • D Cyclic performance of Li/G-LGPS-LCO (LiNbO3 coated) battery.
  • FIGS 36A-36B (A) Voltage profiles of LGPS decomposition at different effective modules (Keff). (B) Reduction reaction pathways corresponding to different Keff and the products in different phase equilibria within each voltage range. All decomposition products here are the ground state phases within each voltage range.
  • FIGS 37A-37F XPS measurement of Ge and P for anode-LGPS-anode symmetric batteries with the X-ray beam focused on (A) the center part LGPS away from the interface to Li/G and (B) the interface between Li/G and LGPS in Li/G-LGPS-G/Li cell under 100 MPa after 12 hours cycle at 0.25 mA cm -2 ; (C) the interface between Li and LGPS in Li-LGPS-Li symmetric battery under 100 MPa after 10 hours cycles at 0.25 mA cm -2 (failed); (D) The Li/G-LGPS interface after rate test at 2 mA cm -2 under 100 MPa and (E) 10 mA cm -2 under 250 MPa; (F) The Li/G-LGPS interface at 2 mA cm -2 under 3 MPa.
  • Figures 38 XRDs of graphite and the mixture of Li and graphite after heating under 500°C for 36 h.
  • Figures 39A-39C SEM images of (A) graphite particles; the surface (B) and cross section (C) of graphite film after applying high pressure.
  • FIGS 41A-44B Comparison of SEM images of Li/G anode before (A) and after (B) long-term cycling in Figure 34(A).
  • FIGS 42A-42C (A) Rate test of Li/G-LGPS-G/Li symmetric battery. When the pre-cycling time is reduced to 5 cycles at 0.25 mA cm -2 , the battery“fails” at 6 mA cm -2 or 7 mA cm -2 , however, when the current density is set back to 0.25 mA cm -2 , it always comes back normal without significant overpotential increase. (B) Enlarged Figure 34(E2), battery cycled at 10 mA cm -2 plotted in a smaller voltage scale (B1) or time scale (B2). (C) SEM images of Li/Graphite composite after testing showing in B with different area and magnification. No lithium dendrite was observed. A clear 3D structure showing this is in Figure 42(C2).
  • FIGS 43A-43B (A) cycling profiles of LCO-LGPS-Li/G batteries in Figure 35D. (B) Cyclic performance based on Li anode. Both batteries were tested at current density of 0.1 C at 25°C.
  • Figures 44A-44B Bader charge analysis from DFT simulations.
  • Figures 45A-45D (A) Comparison of CV curves of Li/G-LGPS-LGPS/C battery tested under 3 and 100 MPa;
  • D Model used in impedance fitting.
  • Rbulk stands for the ionic diffusion resistance and Rct represents the charge transfer resistance. All EIS data are fitted with Z-view.
  • FIGS 46A-46G A CV test of Swagelok battery after they are pressed with 1T, 3T, 6T and pressurized cell initially pressed with 6T.10 % carbon is added in the cathode. The voltage range is set from open circuit to 9.8 V.
  • B The CV scans in (A) plotted in a magnified voltage and current ranges.
  • C In-situ impedance tests during CV scans for batteries shown in (A).
  • D Synchrotron XRD of pressurized cells after no electrochemical process (black), CV scan to 3.2V, 7.5V and 9.8V. All CVs were followed by a voltage holding at the same high cutoff voltages for 10 hours and then discharged back to 2.5V.
  • Green line Synchrotron XRD of LGPS tested in liquid electrolyte after CV scan to 3.2V and held for 10 hours.
  • F Strain versus size broadening analysis for LGPS after high voltage hold. Dots are the broadening of different peaks in 7.5V SXRD measurement, with the corresponding XRD peaks shown in Figure 52. The angle dependences of size and strain broadenings are represented by dashed lines.
  • G XAS measurement of S (g1) and P (g2) after high voltage CV scan and hold.
  • g3 The simulation of P XAS peak shift after straining in the c-direction.
  • FIGS 47A-47D (A) LGPS decomposition energy (a1), ground state pressure (a2), and ground state capacity versus voltage at different effective modules (Keff). (B) Decomposition reaction pathways at different Keff and the products induced by different phase equilibriums in different voltage ranges. (C,D) XPS measurement of S (c) and P (d) element for pristine LGPS (c1, d1), battery after 3.2 V CV scan in liquid electrolyte (c2, d2), pressurized cell after 3.2 V CV scan (c3, d3) and pressurized cell after 9.8 V CV scan (c4, d4). Each CV scan is followed by a 10 hour hold at the high cutoff voltage. Figures 48A-48E.
  • the cyclability of the batteries is represented in (B1), (B2) and (B3) for LCO, LNMO and LCMO, respectively.
  • LCO and LNMO are charged and discharged at 0.3C
  • LCMO is charged at 0.3 C and discharged at 0.1 C. All batteries are tested at room temperature, in the pressurize cell initially pressed with 6T and activate materials are coated with LiNbO3, as shown in Figure 54.
  • C,D XPS measurement of LCO, LNMO, LCMO-LGPS before and after 5 cycles.
  • E XAS measurement of LCO, LNMO, LCMO- LGPS before (E1) and after (E2) 5 cycles for element S.
  • FIGS 49A-49G Pseudo phase simulations of the interface between LGPS and (A) LNO, (B) LCO, (C) LCMO, (D) LNMO. Plots depict the reaction energy of the interface versus the atomic fraction of the non-LGPS phase consumed. The value of the atomic fraction that has the most severe decomposition energy is defined to be ? ? .
  • E-G Mechanically-induced metastability plots for the LGPS-LNO interphase (the set of products that result from the decomposition in Figure 49A).
  • E Energy over hull of the interphase show significant response to mechanical constriction.
  • FIGS 50A-50C (A) Galvanostatic charge and discharge profiles for all-solid-state batteries using LCO and LCMO as cathode and graphite coated lithium metal as anode, with cut-off voltage from 2.6- 4.5 V(LCO) and 2.6- (6-9) V (LCMO).The batteries are charged at 0.3C and discharged at 0.1C. Cycling performance of LCMO lithium metal battery using (B) 1M LiPF6 in EC/DMC and (C) constrained LGPS as electrolyte, with cut-off voltage from 2.5-5.5V with charge rate of 0.3C and discharge rate of 0.1 C.
  • Figure 51 Pellet thickness change in response of force applied.
  • the original thickness of pellet is 756 ⁇ m
  • the weight of the pellet is 0.14 g
  • the area of the pellet is 1.266 cm 2
  • the compressed thickness of the pellet is 250 ⁇ m.
  • the calculated density is 2.1 g/cm 3 , which is close to the theoretical density of LGPS of 2 g/cm 3 .
  • Figures 52A-52F (A)-(F) Synchrotron XRD peaks of batteries at different 2q angles, showing the broadening of XRD peak after high-voltage CV scan and hold. The pressurized cell after 3.2V CV scan and hold doesn’t show XRD broadening.
  • Figure 53 (top) Illustration of decomposition front propagation. Decomposed phases are marked with ?...?. Such propagation is seen to require tangential ionic conduction. (bottom) Energy landscape for reaction coordinates. The final result is a shift in Gibbs energy by D?, which is positive or negative based on equation 2. Even when D? is negative (reaction is thermodynamically favorable), the presence of a sufficient overpotential due to tangential currents can significantly reduce the front’s propagation rate.
  • Figure 54 STEM image and EDS maps of LiNbO3 coated LCO.
  • Figure 56 XAS measurement of LCO, LNMO, LCMO-LGPS before (represented as p) and after (represented as 5c) 5 cycles for element P.
  • Figures 57A-57B (A) Charge and (B) discharge profiles of LCO all-solid-state batteries using LGPS as electrolyte tested with Swagelok, Al pressurized cell, and Stainless steel (SS) pressurized cell with voltage cut-off between 3V-4.15V. Swagelok applied almost no pressure; Al cell is soft compared with Stainless steel and which applied low constrain while stainless steel applied the strongest constant constrain during battery test.
  • the invention provides rechargeable batteries including a solid state electrolyte (SSE) containing an alkali metal and a sulfide disposed between two electrodes.
  • the solid state electrolytes may have a core-shell morphology, imparting increased stability under voltage cycling conditions.
  • These batteries of the invention are advantageous as they may be all-solid-state batteries, e.g., no liquid electrolytes are necessary, and can achieve higher voltages with minimal electrolyte degradation.
  • Core-shell morphologies in which a core of ceramic-sulfide solid-electrolyte is encased in a rigid amorphous shell have been shown to improve the stability window.
  • the strain stabilization mechanism is not limited to the materials level but can also be applied on the battery cell level through external stress or volume constriction.
  • the strain provided by the core-shell structure stabilizes the solid electrolyte through a local energy barrier, which prevents the global decomposition from happening.
  • Such stabilization effect provided by local energy barrier can also be created by applying an external stress or volume constriction from the battery cell, where up to 5.7 V voltage stability window on LGPS can be obtained as shown in Figures 1A-1B. Higher voltage stability window beyond 5.7 V can be expected with higher pressure or volume constriction in the battery cell design based on this technology.
  • lithium dendrites form when the applied current density is higher than a critical value.
  • the critical current density is often reported as 1-2 mA cm -2 at an external pressure of around 10 MPa.
  • a decomposition pathway of the solid state electrolyte, e.g., LGPS, at the anode interface is modified by mechanical constriction, and the growth of lithium dendrite is inhibited, leading to excellent rate and cycling performances. No short-circuit or lithium dendrite formation is observed after the batteries are cycled at a current density up to 10 mA cm -2 .
  • a rechargeable battery of the invention includes a solid electrolyte material and an alkali metal atom incorporated within the solid electrolyte material.
  • solid state electrolytes for use in batteries of the invention may have a core-shell morphology, with the core and shell typically having different atomic compositions.
  • Suitable solid state electrolyte materials include sulfide solid electrolytes, e.g., SixPySz, e.g., SiP2S12 such as Li10SiP2S12, or b/g-PS4.
  • solid state electrolytes include, but are not limited to, germanium solid electrolytes, e.g., GeaPbSc, e.g., GeP2S12 such as Li10GeP2S12, tin solid electrolytes, e.g., SndPeSf, e.g., SnP2S12, iodine solid electrolytes, e.g., P2S8I crystals, glass electrolytes, e.g., alkali metal-sulfide-P2S5 electrolytes or alkali metal-sulfide-P2S5- alkali metal-halide electrolytes, or glass- ceramic electrolytes, e.g., alkali metal-PgSh-i electrolytes.
  • Another material includes germanium solid electrolytes, e.g., GeaPbSc, e.g., GeP2S12 such as Li10GeP2S12, tin solid electrolytes, e.
  • Solid state electrolyte materials are known in the art.
  • the solid state electrolyte material may be in various forms, such as a powder, particle, or solid sheet.
  • An exemplary form is a powder.
  • Alkali metals useful for the solid state electrolytes for use in batteries of the invention include Li, Na, K, Rb, and Cs, e.g., Li.
  • Li-containing solid electrolytes include, but are not limited to, lithium glasses, e.g., xLi2S ⁇ (1-x)P2S5, e.g., 2Li2S ⁇ P2S5, and xLi2S ⁇ (1-x)P2S5–LiI, and lithium glass- ceramic electrolytes, e.g., Li7P3S11-z. Electrode Materials
  • Electrode materials can be chosen to have optimum properties for ion transport.
  • Electrodes for use in a solid state electrolyte battery include metals, e.g., transition metals, e.g., Au, alkali metals, e.g., Li, or crystalline compounds, e.g., lithium titanate such as Li4Ti5O12 (LTO).
  • An anode may also include a graphite composite, e.g., lithiated graphite.
  • Other materials for use as electrodes in solid state electrolyte batteries are known in the art.
  • the electrodes may be a solid piece of the material, or alternatively, may be deposited on an appropriate substrate, e.g., a fluoropolymer or carbon.
  • liquefied polytetrafluoroethylene has been used as the binder when making solutions of electrode materials for deposition onto a substrate.
  • binders are known in the art.
  • the electrode material can be used without any additives.
  • the electrode material may have additives to enhance its physical and/or ion conducting properties.
  • the electrode materials may have an additive that modifies the surface area exposed to the solid electrolyte, such as carbon.
  • Other additives are known in the art.
  • High voltage cathodes of 4 volt LiCoO2 (LCO, shown in Figures 2A-2B) and 4.8V LiNi0.5Mn1.5O4 (LNMO, shown in Figures 3A-3B) are demonstrated to run well in all-solid-state batteries of the invention.
  • Higher voltage cathodes such as the 5.0V Li2CoPO4F, 5.2V LiNiPO4, 5.3V Li2Ni(PO4)F, and 6V LiMnF4 and LiFeF4 may also be used as electrode materials in all-solid-state batteries of the invention. Voltage stability windows beyond 5.7 V, e.g., up to 8 or 10 V or even higher, may be achieved.
  • Another cathode is LiCo0.5Mn1.5O4 (LCMO). Exemplary cathode materials are listed in Table 1, with the calculated stability of the electrodes in Table 1 shown in Figure 4.
  • Table 1 High voltage (greater than 6 V) electrode candidates with individual Materials Project Identifiers.
  • Li4Mn4F16 mp-776813 182.
  • Li2Fe2F8 mp-776813 182.
  • Li4Fe4F16
  • Li6Fe2F12 Li6Fe2F12
  • Li4Fe4F16 Li4Fe4F16
  • Li4Fe2F10 Li4Fe2F10
  • Li4Fe4F16
  • Li4Ge2F12 Li4Ge2F12
  • the electrode materials may further include a coating on their surface to act as an interfacial layer between the base electrode material and the solid state electrolyte.
  • the coatings are configured to improve the interface stability between the electrode, e.g., the cathode, and the solid electrolyte for superior cycling performance.
  • coating materials for electrodes of the invention include, but are not limited to graphite, LiNbO3, AlF3, MgF2, Al2O3, and SiO2, in particular LiNbO3 or graphite.
  • coating materials for anodes with both the required chemical and electrochemical stability. These are generally applicable for LGPS. Table 2 provides the predicted effective coating materials.
  • Zr1Ru1 mp-214 Zr1Zn1: mp-570276 Zr1Zn1Cu2: mp-11366 Zr1Zn1Ni4: mp-11533 Zr1Zn1Rh2: mp-977582 Zr2Be2Si2: mp-10200 Zr2Si2: mp-11322 Zr2Ti2As2: mp-30147 Zr2V2Si2: mp-5541 Zr3Cu4Ge2: mp-15985 Zr3Si2Cu4: mp-7930 Zr4Co4P4: mp-8418 Zr4Mn4P4: mp-20147 Zr4Si4: mp-893
  • Strain stabilization mechanism for enhancing electrolyte stability is not limited to the materials level but can also be applied on the battery cell level through external stress or volume constriction.
  • the external stress is a volumetric constraint applied to all or a portion, e.g., the solid state electrolyte, of the rechargeable battery, e.g., delivered by a mechanical press.
  • the external stress can be applied by a housing, e.g., made of metal.
  • the volumetric constraint can be from about 70 MPa to about 1,000 MPa, e.g., about 70 MPa to about 150 MPa, about 100 MPa to about 300 MPa, about 200 MPa to about 400 MPa, about 300 MPa to about 500 MPa, about 400 MPa to about 600 MPa, about 500 MPa to about 700 MPa, about 600 MPa to about 800 MPa, about 700 MPa to about 900 MPa, or about 800 MPa to about 1,000 MPa, e.g., about 70 MPa, about 75 MPa, about 80 MPa, about 85 MPa, about 90 MPa, about 95 MPa, about 100 MPa, about 150 MPa, about 200 MPa, about 250 MPa, about 300 MPa, about 350 MPa, about 400 MPa, about 450 MPa, about 500 MPa, about 550 MPa, about 600 MPa, about 650 MPa, about 700 MPa, about 750 MPa, about 800 MPa about 850 MPa, about 900 MPa, about 950 MPa, or about 1,000 MPa,
  • the solid state electrolyte may also be compressed prior to inclusion in the battery.
  • the solid state electrolyte may be compressed with a force between about 70 MPa to about 1,000 MPa, e.g., about 70 MPa to about 150 MPa, about 100 MPa to about 300 MPa, about 200 MPa to about 400 MPa, about 300 MPa to about 500 MPa, about 400 MPa to about 600 MPa, about 500 MPa to about 700 MPa, about 600 MPa to about 800 MPa, about 700 MPa to about 900 MPa, or about 800 MPa to about 1,000 MPa, e.g., about 70 MPa, about 75 MPa, about 80 MPa, about 85 MPa, about 90 MPa, about 95 MPa, about 100 MPa, about 150 MPa, about 200 MPa, about 250 MPa, about 300 MPa, about 350 MPa, about 400 MPa, about 450 MPa, about 500 MPa, about 550 MPa, about 600 MPa, about 650 MPa, about 700 MPa, about 750 MPa, about
  • the solid state electrolyte can then be employed in a battery.
  • a battery may also be subjected to external stress to enforce a mechanical constriction on the solid state electrolyte, e.g., at the microstructure level, i.e., to provide an isovolumetric constraint.
  • the mechanical constriction on the solid state electrolyte may be from 1 to 100 GPa, e.g., 5 to 50 GPa, such as about 15 GPa.
  • the external stress required to maintain the mechanical constriction may be from about 1 MPa to about 1,000 MPa, e.g., about 1 MPa to about 50 MPa, about 1 MPa to about 250 MPa, about 3 MPa to about 30 MPa, about 30 MPa to about 50 MPa, about 70 MPa to about 150 MPa, about 100 MPa to about 300 MPa, about 200 MPa to about 400 MPa, about 300 MPa to about 500 MPa, about 400 MPa to about 600 MPa, about 500 MPa to about 700 MPa, about 600 MPa to about 800 MPa, about 700 MPa to about 900 MPa, or about 800 MPa to about 1,000 MPa, e.g., about 70 MPa, about 75 MPa, about 80 MPa, about 85 MPa, about 90 MPa, about 95 MPa, about 100 MPa, about 150 MPa, about 200 MPa, about 250 MPa, about 300 MPa, about 350 MPa, about 400 MPa, about 450 MPa, about 500 MPa, about 550 MPa, about 600 MPa
  • the external stress employed may change depending on the voltage of the battery. For example, a battery operating at 6V may employ an external stress of about 3 MPa to about 30 MPa, and a battery operating at 10V may employ an external stress of about 200 MPa.
  • the invention also provides a method of producing a battery using compression of the solid state electrolyte prior to inclusion in the battery, e.g., with subsequent application of external stress.
  • Batteries of the invention may be charged and discharged for a desired number of cycles, e.g., 1 to 10,000 or more.
  • batteries may be cycled 10 to 750 times or at least 50, 100, 200, 300, 400, 500, 600, 700, 800, 900, 1,000, 1,500, 2,000, 3,000, 4,000, or 5,000 times.
  • the voltage of the battery ranges from about 1 to about 20V, e.g., about 1-10V, about 5-10V, or about 5- 8V.
  • Batteries of the invention may also be cycled at any appropriate current density e.g., 1 mA cm -2 to 20 mA cm -2 , e.g., about 1-10 mA cm -2 , about 3-10 mA cm -2 , or about 5-10 mA cm-2.
  • the cyclic voltammograms (CV) of Li/LGPS /LGPS+C were measured under different pressures between open circuit voltage (OCV) to 6 V at a scan rate of 0.1mVs -1 on a Solartron electrochemical potentiostat (1470E), using lithium (coated by Li2HPO4) as reference electrode.
  • OCV open circuit voltage
  • Li2HPO4 Solartron electrochemical potentiostat
  • a liquid battery using LGPS/C thin film as cathode, lithium as anode and, 1 M LiPF6 in EC/DMC as electrolyte was also assembled for comparison.
  • the ratio of LGPS to C is 10:1 in both solid and liquid CV tests.
  • the cathode and anode thin films used in all-solid-state battery were prepared by mixing
  • the ratios of active materials/LGPS/C are 30/60/10, 70/27/3, 70/30/0 for LTO, LCO and LNMO thin film electrodes, respectively.
  • This mixture of powder was then hand-grinded in a mortar for 30 minutes and rolled into a thin film inside an argon-filled glove box with 3% PTFE added.
  • Solid electrolytes used in all-solid-state Li ion batteries were prepared by mixing LGPS and PTFE with a weight ratio of 97:3, then hand-grinding the mixed powder in a mortar for 30 minutes and finally rolling it into a thin film inside an argon-filled glove box.
  • any mechanical constraint will require that decomposition induce strain in the surrounding neighborhood.
  • a constraining system could be either materials- level (i.e. a core-shell microstructure) or systems-level (i.e. a pressurized battery cell) or a combination of the two.
  • this mechanical system can only induce a finite strain before fracturing. The energy needed to fracture the system is denoted
  • the particle begins as pristine LGPS with an unfractured constraint mechanism 2. As the particle begins to decompose the constraint mechanism requires an increase in The strain Gibbs is assumed to be a function of ? ? that goes to zero as goes to zero
  • ? ??? is the undeformed volume, ? is the strain tensor relative to the undeformed state and ? is the stress tensor corresponding to ?.
  • the solid-like work can be separated into one term that only includes compression and one term that only includes deformation.
  • equation 3 reduces to
  • the fluid term - indicates the work needed to compress the reference volume (i.e., change ? ? ) in the presence of a stress tensor ? and the solid term represents the work needed to deform the new reference state Considering this, the full energy differential is given by
  • equation 6 gives the differential form of of Figures 5A-5B in terms of t he chemical terms and the strain term
  • the first case is that of a LGPS particle that decomposes hydrostatically and is a mean field approximation.
  • the fraction of decomposed LGPS is assumed to be uniform throughout the particle
  • the second limiting case is that of spherically symmetric nucleation, where LGPS is completely decomposed within a spherical region of radius and pristine outside this region
  • the local stress ? experienced by a subsection of an LGPS particle is directly a function of the decomposition profile as well as the mechanical properties of the particle and, if applied, the mechanically constraining system.
  • the local stress is said to be compressive and equal everywhere within the particle In the mean field
  • the particle is known to be at least metastable with total stability being determined by the magnitude of The relationship between the pressure and decomposed fraction was shown in r to be, in this limit, .
  • the reference volume is the volume in the unconstrained system .
  • Equation 9 is solved for in Figure 5 for the case of a core-shell constriction mechanism with a core comprised of either LGPS or oxygen-doped LGPSO and a shell of an arbitrary rigid material.
  • the effective bulk modulus is given by is the compressibility of the LGPS material and is a parameter that represents the ability
  • the maximally localized (i.e. highest local pressure) decomposition mechanism is that of spherical nucleation as shown in Figure 6.
  • an LGPS particle of outer radius ? ? undergoes a decomposition at its center.
  • the decomposed region corresponds to the material that was initially within a radius of ? ? .
  • the new reference state is of higher volume than the pristine state as the material has decomposed to a larger volume given by
  • the decomposed fraction is no-longer a constant in the particle as it was in the hydrostatic case. Instead, for all material that was initially (prior to decomposition) within the region for all
  • the boundary conditions are:
  • the displacement vector is known to be of the form ? , where the vector notation has been removed as displacement is only a function of distance from the center.
  • the strain Gibbs for a compressed sphere under condition 2 defining gives the compressive term ? with no deviatoric components.
  • a hollow pressurized sphere at the onset of decay lim has both a compressive and deviatoric component that combine to where ?? is the shear modulus of the pristine material. Combining these terms leads to the nucleated equivalent of equation 8.
  • the gray and purple lines reflect the no-shell and perfect-shell limits of the hydrostatic model, whereas the blue and red lines represent equation 10 for typical Poisson values. It is seen that, in general, the nucleation model provides a steeper strain Gibbs than the hydrostatic model due to the higher pressures involved. Intuitively, a smaller Poisson’s ratio (harder to compress) improves the stability of the nucleation limit.
  • Electrolytes either liquid or solid, are likely to react with electrodes where the electrode potential is outside of the electrolyte stability window. To address this, it is suggested that electrolytes be chosen such that they form a passivating solid-electrolyte-interface (SEI) that is at least kinetically stable at the electrode potential.
  • SEI solid-electrolyte-interface
  • Many works on the topic of improving sulfide electrolytes have speculated that by forming electronically insulating layers on the surface of sulfide electrolytes such passivation layers can be formed. In this section, we discuss the role of such passivation layers and provide a quantitative analysis of the mechanism by which we believe an electronically insulating surface layer improves stability.
  • thermodynamic equilibrium state is given for the most basic battery half-cell model.
  • the voltage of the lithium metal is defined to be the zero point.
  • the differential Gibbs energy can be written as equation 12 (superscripts ?, ? differentiate the anode from the cathode).
  • Figure 8B depicts the expected equilibrium state in the case of a solid-electrolyte cathode, where the cathode material is imbedded in a matrix of solid-electrolyte.
  • the lower (i.e. more- negative) chemical potential of the cathode material relative to the electrolyte causes charge separation that results in an interface voltage ? ? .
  • the equilibrium points now include the anode (a), cathode (c) and the solid- electrolyte (SE):
  • equation 14 leads to the condition that the lithium metal potential remains constant throughout the cell.
  • FIG. 8C A speculated mechanism for passivation layer stabilization of sulfide electrolytes is depicted in Figure 8C.
  • the solid-electrolyte is coated in an electronically insulating material. Since the external circuitry does not directly contact the solid-electrolyte and there is no electron conducting pathway, the number of electrons within the solid-electrolyte is fixed. Hence the Fermi energy cannot equilibrate via electron flow. The speculation is that this effect could be utilized to allow a deviation of the lithium metal potential within the solid-electrolyte relative to the electrodes, leading to a wider operational voltage window.
  • the band diagrams of Figure 8C illustrate how the electron
  • electrochemical potential can experience a local maximum (or minimum) in the solid-electrolyte due to a lack of electron conduction. This local maximum (or minimum) is carried over to the lithium metal potential.
  • Constraints 1 and 2 represent the tethering of the electron and lithium density in the case of an insulated particle.
  • the Fermi level of the solid-electrolyte is not fixed by an external voltage. The result is that by lowering the number of atoms within the solid- electrolyte by extracting lithium ions, and hence increasing the number of electrons per atom within the insulated region, the number of electrons per atom and the Fermi level increase. In effect, this represents the conduction of electrons by way of lithium-holes.
  • Solving equation 15 for the equilibrium points given the above constraints lead to those of equation 14 between the anode/cathode as well as the following relation between the anode and solid-electrolyte.
  • the total voltage experienced within the SE can be represented as where is the voltage in the absence of lithium extraction from the SE (the original voltage as depicted in Figure 8C) and ? ? is the voltage that results from the charge separation of lithium extraction.
  • the system begins with a charge neutral solid-electrolyte at voltage ? ??
  • ultra-LGPS decomposition of ultra-LGPS was largely suppressed, manifested by only one minor oxidation peak at a higher voltage (3V) during charging, and almost no reduction peak during discharging (Figure 15B).
  • the higher stability of ultra-LGPS is also confirmed by the sensitive electrochemical impedance spectra (EIS) before and after CV tests ( Figures 15C, 15D).
  • EIS shows a typical Nyquist plot of battery-like behavior with charge-transfer semicircles in the medium frequency and a diffusion line in the low frequency.
  • FIG. 15E shows the charge-discharge profiles of LGPS (LTO+LGPS+C/Glass fiber separator/Li) cycled at 0.5C in the voltage range of 1.0 - 2.2 V.
  • the plateau length decreases from cycle 1 to cycle 70 by almost 85.7%, indicating a large decay of the cathode.
  • ultra-LGPS LTO+ultra-LGPS+C/Glass fiber separator/Li
  • Figure 15F shows the same flat voltage plateau remaining almost unchanged after 70 cycles.
  • This increase in cathode stability is further confirmed by the cyclic capacity curves ( Figures 15G and 15H).
  • the specific charge and discharge capacities decrease from ⁇ 159 mAh/g to ⁇ 27 mAh/g, and ⁇ 170mAh/g to ⁇ 28 mAh/g, respectively, after 70 cycle.
  • ultra-LGPS demonstrates a much better cyclic stability than its LGPS counterpart. After 70 cycles the discharge capacity is still as high as 160 mAh/g, with only roughly 5% of capacity loss.
  • ultra-LGPS has, in practice, improved stability over LGPS in the cases of both LGPS oxidation and reduction. Additionally, the Coulombic efficiency of ultra-LGPS is also higher than that of LGPS, indicating an improved efficiency of charge transfer in the system, and less charge participation in unwanted side reactions.
  • a transit layer with multiple small dark particles exists at the cathode/separator interface (hereafter“LTO/LGPS primary interface), as manifested in the TEM bright-field (BF) images ( Figure 19B, Figure 20) and STEM dark-field (DF) images ( Figure 19D, Figure 20).
  • the particles within the transit layer of STEM DF images show bright contrast, indicating the accumulation of heavy elements.
  • STEM EELS electron energy loss spectroscopy
  • LTO/LGPS secondary interface LTO/LGPS interfaces
  • sulfur-rich particles exist at both primary and secondary LTO/LGPS interfaces in LGPS half-cells after 1 charge-discharge cycle.
  • Figures 23A-23F show the microstructural and compositional (S)TEM studies for ultra-LGPS half-cells.
  • the primary LTO/ultra-LGPS interface after 1 charge-discharge cycle was characterized by TEM BF image ( Figure 23A).
  • a smooth interface was observed between the ultra- LGPS separating layer and the composite cathode layer ( Figure 23B).
  • the primary LTO/ultra-LGPS interface is clean and uniform, showing no transit layer or dark particles.
  • the secondary LTO/ultra- LGPS interfaces were also investigated for comparison by STEM DF image, EDS line-scan and EDS mapping ( Figures 23C-23E).
  • Results show that the atomic percentage of sulfur continuously decreases, as the STEM EDS line-scan goes from inner ultra-LGPS particle to secondary LTO/ultra- LGPS interface, and finally into LTO+C composite region ( Figure 23D and Figures 24A, 24B).
  • the sulfur-deficient-shell feature of ultra-LGPS particles is maintained after cycling, and no sulfur-rich transit layer is formed at the LTO/ultra-LGPS secondary interface.
  • STEM EDS quantitative analyses Figure 23F show that the atomic percentage of sulfur inside ultra-LGPS particle is as high as ⁇ 38%, while that of secondary LTO/ultra-LGPS interface is as low as 8%.
  • LGPS powder was purchased from MSE Supplies company. Ultra-LGPS was synthesized by soaking LGPS powder into organic electrolytes, such as dimethyl carbonate (DMC) and diethyl carbonate (DEC), and then sonicated for 70h in Q125 Sonicator from Qsonica company, a microprocessor based, programmable ultrasonic processor
  • DMC dimethyl carbonate
  • DEC diethyl carbonate
  • the cyclic voltammograms (CV) of Li/LGPS/LGPS+C/Ta and Li/ultra-LGPS/ultra-LGPS/Ta cells were measured between 0.5 to 5 V at a scan rate of 0.1mVs -1 on a Solartron electrochemical potentiostat (1470E), using lithium as reference electrode.
  • the electrochemical impedance spectrums of Li/LGPS/LGPS+C/Ta and Li/ultra-LGPS/ultra-LGPS/Ta cells were measured at room temperature both before and after CV tests, by applying a 50 mV amplitude AC potential in a frequency range of 1 MHz to 0.1 Hz.
  • the composite cathode used were prepared by mixing LTO, (ultra-)LGPS, polyvinylidene fluoride (PVDF) and carbon black with a weight ratio of 30:60:5:5. This mixture of powders was then hand-grinded in a mortar for 30 minutes and rolled into a thin film inside an argon- filled glove box. SEs were prepared by mixing (ultra-)LGPS and PVDF with a weight ratio of 95:5, then hand-grinding the mixed powder in a mortar for 30 minutes and finally rolling it into a thin film inside an argon-filled glove box.
  • the prepared composite cathode thin film, (ultra-)LGPS thin film, and Li metal foil were used as cathode, solid electrolyte, and the counter electrode, respectively.
  • the thin films of composite cathode and (ultra-)LGPS were cold-pressed together before assembling into the battery.
  • a piece of glass fiber separator was inserted between (ultra-)LGPS thin film and Li metal foil to avoid interfacial reaction between these two phases. Only 1 drop of 1 M LiPF6 in ethylene carbonate (EC) and dimethyl carbonate (DMC) solution (1:1) was carefully applied onto the glass fiber to allow lithium ion conduction through the separator.
  • Swagelok- type cells were assembled inside an argon-filled glove box.
  • Assembling process of an (ultra-)LGPS battery is the same with that of an (ultra-)LGPS solid-state battery, except that the (ultra-)LGPS SE layer is removed.
  • the charge/discharge behavior was tested using an ArbinBT2000 workstation (Arbin Instruments, TX, USA) at room temperature. The specific capacity was calculated based on the amount of LTO (30 wt%) in the cathode film.
  • FIB sample preparation For FIB sample preparation, the cold-pressed thin film of composite cathode and (ultra-)LGPS after 1 charge-discharge cycle in (ultra)LGPS solid-state battery was taken out inside an argon-filled glove box. It was then mounted onto a SEM stub and sealed into a plastic bag inside the same glove box. FIB sample preparation was conducted on an FEI Helios 660 dual-beam system. The prepared FIB sample was then immediately transferred into JOEL 2010F for TEM and STEM EDS/EELS characterization. Density functional theory calculations
  • the key performance metrics for solid-electrolytes are stability and ionic conductivity.
  • two very promising families of solid-electrolytes are garnet-type oxides and ceramic sulfides. These families are represented, respectively, by the high-performance electrolytes of LLZO oxide and LSPS sulfide. Oxides tend to maintain good stability in a wide range of voltages but often have lower ionic conductivity Conversely, the sulfides can reach excellent ionic conductivitie but tend to decompose when exposed to the conditions needed for battery operation.
  • Instabilities in solid-electrolytes can arise from either intrinsic material-level bulk decompositions or surface/interfacial reactions when in contact with other materials. At the materials-level, solid- electrolytes tend to be chemically stable (i.e. minimal spontaneous decomposition) but are sensitive to electrochemical reactions with the lithium ion reservoir formed by a battery cell.
  • the voltage stability window defines the range of the lithium chemical potential within which the solid-electrolyte will not electrochemically decompose.
  • the lower limit of the voltage window represents the onset of reduction, or the consumption of lithium ions and the corresponding electrons, whereas the upper limit represents the onset of oxidation, or the production of lithium ions and electrons.
  • the voltage window affects the bulk of any solid-electrolyte particle as the applied voltage is experienced throughout. While interfacial reactions occur between the solid-electrolyte and a second‘coating’ material at the point of contact, these reactions can either be two-bodied chemical reactions, where only the solid- electrolyte and the coating material are reactants, or three-bodied electrochemical reactions, in which the solid-electrolyte, coating material and the lithium ion reservoir all participate. The two types of reactions are state-of-charge or voltage independent and dependent, respectively, as determined by the participation of the lithium ion reservoir.
  • dimensionality of the problem is governed by the number of elements. For example, calculating the interfacial chemical stability of LSPS and LCO would require a 6-dimensional hull corresponding to the set of elements ⁇ Li, Si, P, S, Co, O ⁇ . The electrochemical stability of this interface is calculated with the system open to lithium, so that lithium is removed from the set and the required hull becomes 5-dimensional ( ⁇ Si, P, S, Co, O ⁇ ).
  • This hull is the same hull that must be calculated for the interface with LFPO and includes, as a subset, the 5- dimensional hull needed for the evaluation of iron-sulfide (???).
  • the minimum number of elemental sets that spans the entirety of the materials were determined ( Figure 25A). Then for each elemental set, only one hull is needed to evaluate all of materials that can be constructed using those elements. This approach reduces the total number of hulls needed from 67,062 (one per material) to 11,935 (one per elemental set).
  • Figure 25A few hulls with a dimensionality below 7 were needed. Those compounds that would otherwise require a low dimensional hull are solved as a subset of a larger element set. Additionally, the number of required 7 and 8 dimensional hulls are largely reduced due to multiple phases of the same compositional space requiring the same hull.
  • the second schema used to minimize computational cost was a binary search algorithm for determining the pseudo-binary once a hull was calculated.
  • the pseudo-binary approach is illustrated in Figure 25B. Since decomposition at an interface between two materials can consume an arbitrary amount of each material, the fraction of one of the two materials (? in equation 1) consumed can vary from 0-1.
  • the pseudo-binary is a computational approach that determines for which value of ? the
  • Equation 1 is the most kinetically driven (e.g. when is the decomposition energy the most severe).
  • the RHS of equation 1 represents the fraction ( ⁇ ? ? ⁇ ) of each of the thermodynamically favored decay products and defines the convex hull for a given ? in terms of the products’ Gibbs energies The total decomposition energy accompanying
  • equation 1 is:
  • the most kinetically driven reaction between LSPS and the coating material is the one that maximizes the magnitude (i.e. most negative) of equation 2, which defines the parameter ? ? .
  • This maximum decomposition energy is the result of two factors. The first, denoted , is the portion
  • Figure 25C indicates that chemical stability is best for those compounds that contain large anions such as sulfur, selenium and iodine.
  • Figures 26A and 26C indicate that there is reduced correlation between elemental species and at low and high voltages, respectively. This suggests that at these voltage extremes, the interfacial decomposition is dominated by intrinsic materials-level reduction/oxidation rather than interfacial effects lithium (Figure 26B) positive correlation (higher instability) is seen for most elements with the notable exception of the chalcogen and halogen anion groups, which are negatively correlated.
  • Figure 27A illustrates the impact of applied voltage on the hull energy of a material, in this case LSPS.
  • The“neutral decay” line at 45 ⁇ represents those compounds that have the same hull energy at both voltage extremes and hence aren’t reacting with the lithium ions. Datapoints above [below] this line are increasing [decreasing] in hull energy with respect to voltage and are hence are characteristically oxidative [reductive] in the plotted voltage range.
  • Figure 27B indicates that, in agreement with expectations, most compounds are reduced in the anode voltage range of 0-1.5 V vs. lithium metal. Nitrogen containing compounds are seen to
  • the average hull energy of each anionic class is given in 0.5V steps from 0-5V in Figure 27D.
  • Nitrogen containing compounds are confirmed to be the most stable at 0V with iodine and phosphorous compounds maintaining comparable stability. Phosphorous and iodine surpass nitrogen in average stability for voltages above 0.5V and 1.0V, respectively. At high voltages (>4V), it is seen that fluorine and iodine containing compounds are stable whereas nitrogen containing compounds are the least stable.
  • Figure 28A shows the average instability due to chemical reactions between the anionic classes and LSPS. Sulfur and selenium containing compounds form, on average, the most chemically inert interfaces with LSPS. Conversely, fluorine and oxygen containing compounds are the most reactive. As a general trend, those compound classes that are more unstable in total terms (higher ? ???? ? ? ? ? ) also maintain a higher interfacial contribution relative to the intrinsic material contribution
  • Figure 28B shows the average total electrochemical decomposition energy for the interfaces in 0.5V steps from 0-5V.
  • each anionic class follows a path that appears to be dominated by the materials-level electrochemical stability of LSPS (Figure 27A). This is particularly true in the low voltage ( ⁇ 1V) and high voltage (>4V) regimes, where electrochemical effects will be the most pronounced.
  • the biggest deviations of the interfacial stability from LSPS’s intrinsic stability occur in the region of 1-3V.
  • each anionic class that were determined to be functionally stable or potentially functionally stable are given in Figure 29A (anode range) and Figure 29B (cathode range), where they are both intrinsically stable at the material level and form stable interfaces with LSPS within the prescribed voltage range.
  • nitrogen, phosphorous, and iodine containing compounds have the highest percentage of stable compounds (2-4%), whereas all other classes are below 1%.
  • the cathode range showed much higher percentages with sulfur containing compounds reaching 35%. Iodine and selenium were both above 10%.
  • the mixed powders were annealed at high temperatures (300°C, 400°C, 500°C) to determine the onset temperature of interfacial reactions as well as the reaction products, and to further assess the role of kinetics by comparing these results with the DFT computed thermodynamic reaction products.
  • Figures 30A-30D compares the XRD patterns of such room-temperature and 500°C-annealed powder mixtures.
  • candidate coating materials i.e. SnO2, Li4Ti5O12, SiO2
  • Figures 30C-30D the mixed powder of LCO+LSPS was for comparison
  • Figure 30A The XRD patterns for each individual phase (i.e. SnO2, Li4Ti5O12, LiCoO2, SiO2 and LSPS) at room temperature and 500°C are used as reference ( Figures 31A-31E).
  • the electrochemical stability of typical coating materials is characterized by Cyclic Voltammetry (CV) technique, in which the decomposition of the tested coating material can be manifested by current peaks at certain voltages relevant to Lithium.
  • CV Cyclic Voltammetry
  • Two typical coating materials were used as a demonstration to show good correspondence between our theoretical prediction and experimental observation.
  • the CV test of Li2S ( Figure 30E) shows a relevantly flat region between 0-1.5V, while a large oxidation peak dominates the region of 2-4V.
  • the CV test of SiO2 (Figure 30F) demonstrates net reduction in the region of 0-1.5V, and a neutral region with little decomposition between 2 and 4V.
  • the smallest number of elemental sets that spanned all the materials were determined. To do this, the set of elements in each structure were combined with the elements of LSPS, resulting in a list of element sets with each set’s length equal to the dimensionality of the required hull for that material. This list was ordered based on decreasing length of the set (e.g. ordered in decreasing dimensionality of the required hull). This set was then iterated through and any set that equals to or is a subset of a previous set was removed. The result was the minimum number of elemental sets, in which every material could be described. Chemical decomposition hulls were calculated using the energies and compositions from the MP.
  • the pseudo-binary seeks to find the ratio of LSPS to coating material such that the decomposition energy is the most severe and, hence, is the most kinetically driven.
  • This problem is simplified by using a vector notation to represent a given composition by mapping atomic occupation to a vector element. For example, in the basis of (Li Co O), meaning that there are 1 lithium, 1 cobalt, and 2 oxygen in the unit formula. Using this notation, the decomposition in equation 1 can be written in vector form.
  • equation 5 Using ? to represent a vector and to represent a matrix, equation 5 becomes:
  • Equation 7 allows for the calculation of the derivative of the hull energy with respect to the fraction parameter ?.
  • Equation 7 allows for the calculation of the derivative of the hull energy with respect to the fraction parameter ?.
  • Equations 7 and 8 can then be used to find the slope of the hull energy. If the hull energy is positive, , whereas if it is negative This process is repeated until the upper and lower limits differ by a factor less than the prescribed threshold of 0.01%, which will always be achieved in 14 steps (
  • Equations 5-8 are defined for chemical stability. In the case of electrochemical (lithium open) stability, the free energy is replaced with F where ? is the chemical potential and ? ? is the number of lithium in structure ?. Additionally, lithium composition is not included in the composition vectors of equation 6 to allow for the number of lithium atoms to change.
  • XRD X-ray diffraction
  • XRD tests were performed on Rigaku Miniflex 600 diffractometer, equipped with Cu Ka radiation in the 2-theta range of 10-80°. All XRD sample holders were sealed with Kapton film in Ar-filled glovebox to avoid air exposure during the test.
  • Candidate coating materials Li2S and SiO2
  • carbon black carbon black
  • the powder mixtures were sequentially hand-rolled into a thin film, out of which circular disks (5/16-inch in diameter, ⁇ 1-2 mg loading) were punched out to form the working electrode for Cyclic
  • the Li/graphite anode was designed as shown in Fig.33(D).
  • the protective graphite film was made by mixing graphite powder with PTFE and then covering onto the lithium metal.
  • the three layers of Li/graphite, electrolyte and cathode film were stacked together sequentially, followed by a mechanical press.
  • the pressure was maintained at 100-250 MPa during the battery test. Such pressure helps obtain a good contact between anode and electrolyte based on the conventional wisdom in this field, but more importantly, it serves a mechanical constriction for improved electrochemical stability of solid electrolyte.
  • Scanning electron microscopy (SEM) shows that the graphite particles transform into a dense layer under such high pressure (Fig.39).
  • the as-prepared anode before battery test can be directly observed via SEM and focused ion beam (FIB)-SEM in Fig.33E, 33F).
  • the three layers of Li, graphite and LGPS were clear with close interface contact.
  • Li/graphite (Li/G) anode was tested with anode- LGPS-anode symmetric battery design under 100 MPa external pressure.
  • the comparison of cyclic performance between Li/G-LGPS-G/Li and Li-LGPS-Li batteries is shown in Fig.34A.
  • Li symmetric battery works only for 10 hours at a current density of 0.25 mA cm -2 before failure, while Li/G symmetric battery was still running after 500 hours of cycling with the overpotential increasing slowly to 0.28 V.
  • the stable cyclic performance was repeatable, as shown in Fig.40 from another battery with a slower overpotential increase from 0.13 V to 0.19 V after 300 hours’ cycling, indicating such slight overpotential change varies from battery assembly.
  • Li/G symmetric battery under different external pressures of 100 MPa or 3 MPa as shown in Fig.34C. Same charging and discharging capacities were set for different current densities by changing the working time per cycle.
  • the Li/G symmetric battery can cycle stably from 0.25 mA cm -2 up to 3 mA cm -2 with an overpotential increase from 0.1 V to 0.4 V. It can then cycle back normally to 0.25 mA cm -2 (Fig.34C1). While at 3 MPa, the battery failed during the test at 2 mA cm -2 (Fig.34C2). Note that at the same current density, the overpotential at 100 MPa was only around 63 % of that under 3 MPa.
  • Keff the effective modulus
  • the effective modulus represents the intrinsic bulk modulus of the electrolyte added in parallel with the finite rigidity of the battery system. Accordingly, Keff measures the mechanical constriction that can be realized on the materials level in any single particle, while the external pressure applied on the operation of solid state battery enforced the effectiveness of such constriction on the interface between particles or between electrode and electrolyte layers. This is because exposed surface was the most vulnerable to chemical and electrochemical decompositions, while a close interface contact enforced by external pressure will minimize such surface. Thus, even though the applied pressure was only on the order of 100 MPa, the effective bulk modulus was expected to be much larger.
  • thermodynamic overpotential (?) dominates and favors the ground state decomposition products of Figure 36.
  • ? ? begins to dominate and favors those metastable phases, such as LixGey at high in our computations, which are not shown in Fig.36 as those are all ground state phases in each voltage range.
  • a lithium-graphite composite allows the application of a high external pressure during the test of solid- state batteries with LGPS as electrolyte. This creates a high mechanical constriction on the materials level that contributes to an excellent rate performance of Li/G-LGPS-G/Li symmetric battery. After cycling at high current densities up to 10 mA cm -2 for such solid-state batteries, cycling can still be performed normally at low rates, suggesting that there is no lithium dendrite penetration or short circuit.
  • the reduction pathway of LGPS decomposition under different mechanical constrictions are analyzed by using both experimental XPS measurements and DFT computational simulations. It shows, for the first time, that under proper mechanical constraint, the LGPS reduction follows a different pathway.
  • Graphite thin film is made by mixing active materials with PTFE.
  • All the batteries are assembled using a homemade pressurized cell in an argon-filled glovebox with oxygen and water ⁇ 0.1 ppm.
  • the symmetric battery Li/G-LGPS-G/Li or Li- LGPS-Li was made by cold pressing three layers of Li(/graphite)-LGPS powder- (graphite/)Li together and keep at different pressures during battery tests. The batteries were charged and discharged at different current densities with the total capacity of 0.25mAh cm -2 for each cycle.
  • a LiCoO2 half battery was made by cold pressing Li/graphite composite-LGPS powder-Cathode film using a hydraulic press and keep the pressure at 100-250 MPa.
  • the LiCoO2 were coated with LiNbO3 using sol-gel method.
  • the CV test (Li/G-LGPS-LGPS/C) was conducted on a Solartron 1400 cell test system between OCV to 0.1V with the scan rate of 0.1 mV/s.
  • XRD XRD XRD
  • Example 5 In this work, we focused on how the external application of either high-pressure or isovolumetric conditions can be used to stabilize LGPS at the materials level through the control at the cell-level. This advances beyond the microstructural level mechanical constraints present in previous works, where particle coatings were used to induce metastability. Under proper mechanical conditions, we show that the stability window of LGPS can be widened up to the tool testing upper limit of 9.8 V. Synchrotron X- ray diffraction (XRD) and x-ray absorption spectroscopy (XAS) that measure the structure changes of LGPS before and after high-voltage holding show, for the first time, direct evidence of LGPS straining during these electrochemical processes.
  • XRD X- ray diffraction
  • XAS x-ray absorption spectroscopy
  • thermodynamic and kinetic factors are further considered by comparing density functional theory (DFT) simulations and x-ray photoelectron spectroscopy (XPS) measurements for decomposition analysis beyond the voltage stability window.
  • DFT density functional theory
  • XPS x-ray photoelectron spectroscopy
  • Li4Ti5O12 (LTO) anodes are paired with LiCo0.5Mn1.5O4 (LCMO), LiNi0.5Mn1.5O4 (LNMO) and LiCoO2 (LCO) cathodes to demonstrate the high-voltage stability of constrained LGPS.
  • LCMO LiCo0.5Mn1.5O4
  • LNMO LiNi0.5Mn1.5O4
  • LCO LiCoO2
  • the density of the LGPS pellets after being pre-pressed at 1, 3, and 6T were 62%, 69% and 81%, respectively, of the theoretical density of single crystal LGPS.
  • the morphology of LGPS pellets after pressing is shown in Fig. 51A.
  • Figure 46G shows the P and S XAS peaks of pristine LGPS compared with the ones after CV scan up to 3.2V and 9.8V in liquid or solid-state batteries.
  • 3.2V-L the conditions of no mechanical constraint
  • both P and S show obvious peak shift toward high energy and the shape change, indicating significant global oxidation reaction and rearrangement of local atomic environment in LGPS in the liquid cell.
  • the P and S peaks don’t show any sign of global oxidation in solid state batteries, as no peak shift is observed.
  • FIG. 47A1 shows the energy above the hull, or the magnitude of the decomposition energy.
  • An energy above the hull of 0 eV atom -1 indicates that thermodynamically the LGPS is the ground state product, whereas an elevated value indicates that the LGPS will decay.
  • the region in which the energy above the hull is nearly zero ( ⁇ 50 meV for thermal tolerance) is seen to increase in upper voltage limit from approximately 2.1V to nearly 4V.
  • Figure 47A2 shows the ground state pressure corresponding to the free energy minimization. The pressure is given by where ? corresponds to the fraction volume transformation of LGPS to the products that
  • the application of the mechanical constraint can greatly reduce the speed at which ceramic sulfides decay as depicted in Figure 53.
  • the effective stability the “mechanically-induced kinetic stability”– was sufficiently high as to allow battery operation. For example, if the electrolyte only decays one part per million per charge cycle, then it was sufficiently stable for practical battery designs that only need last thousands of cycles.
  • the proposed mechanism for mechanically-induced kinetic stability is depicted in Figure 53.
  • the third region is the interface, where the mole fraction transitions from 0 to 1.
  • the propagation direction of the decomposition front is controlled by thermodynamic relation of Equation 1. If Equation 1 is satisfied, the front will propagate inwards, preferring the pristine LGPS. Accordingly, the LGPS will not decompose. When Equation 1 is violated, the front will propagate into the LGPS and ultimately consume the particle.
  • Equation 1 the speed with which the front propagates into the pristine LGPS will still be influenced by the application of mechanical constraint. This is illustrated in Figure 53 (bottom).
  • the decomposition front propagates, there must exist ionic currents tangential to the front’s curvature. This requires the presence of an overpotential to accommodate the finite conductivity of the front for each elemental species.
  • the ohmic portion of the overpotential is given by the sum of equation 3, where is the resistivity of the front for each species ? at the pressure (?) that is present at the front, is the characteristic length scale of the decomposed morphology, and ? ? is the ionic current density. Given that ? ? ???
  • strain of a single lattice vector is approximately the strain of the ab-plane of LGPS near the
  • activation energy for Li migration in LGPS is predicted to increase from 230 meV to 590 meV upon constriction by 4%, the rate at which lithium reordering can occur decreases by a factor of:
  • Figure 48 shows the galvanostatic cycling along with their cyclability performance of all-solid-state batteries, using LCO, LNMO and LCMO as cathode, LGPS as a separator and LTO as anode.
  • the battery tests were performed in the pressurized cell, where the cells were initially pressed with 6T then fastened in bolted [quasi]-isovolumetric cell.
  • LCO is the most common and widely used cathode material, included in commercial Li-ion batteries, with a plateau at approximately 4 V against Li + /Li
  • LNMO is considered one of the most promising high voltage cathode materials with a flat operating voltage at 4.7 V versus Li + /Li.
  • FIG. 55 The high rate test of LCO full battery is shown in Figure 55.
  • the charge and discharge curves of LCO and LNMO are depicted in Figs.48A1 and 48B1, respectively.
  • Both batteries show a flat working plateau centered at 2 V (3.5 V vs Li + /Li) for LCO and 2.9 V (4.4 V vs. Li + /Li) for LNMO in the first discharge cycle.
  • both of them exhibit excellent cyclability performance, as can be observed in Figs.48A2 and B2, with a capacity fading of just 9% in the first 360 cycles for LCO and 18% in the first 100 cycles for LNMO. This is an indication that the decomposition or interfacial reaction of the cathode materials with LGPS was not very severe.
  • FIGS 48C1-48D3 show the XPS measured binding energy of electrons in LGPS before and after battery cycles using LCO, LNMO and LCMO as cathodes. Each element can become oxidized either by chemical reaction with the cathode material (chemical oxidation) or the delithiation of the LGPS by the application of a voltage (electrochemical oxidation).
  • XAS measurement shows a pre-edge on the intensity of S element while no pre-edge is found from P ( Figures 48E and 56), given that S, instead P, is bonded with trasition metal, no matter from coating materials or cathode materials. Althought the interface reaction is evaliated by the mechanical constraint, there is still a ceterin amount of side reactions happens from the direct contract between cathode materials and LGPS. More interface reactions occur after battery cycles.
  • Figures 49A-D the atomic fraction of the cathode material (or LNO) is swept from 0 to 1 (representing pure LGPS to pure cathode or LNO). Whichever value of atomic fraction makes the reaction energy the most negative represents the worst-case reaction and is termed ? ? . Table 6 gives these ? ? values for each interface, along with the worst-case reaction energy, the decomposed products, and an additional pseudo-phase that represents the decomposed interface. This pseudo-phase that represents the decomposed interface, also known as the interphase, can be used to calculate how the decomposed interface will further decay as the battery is cycled.
  • Figure 49E-G show the electrochemical stability of the LGPS+LNO interphase.
  • Figures 49B-D show that the chemical reaction energies for LCO, LNMO, and LCMO are 345, 322, and 335 meV atom -1 , respectively.
  • LNO which has a much lower reaction energy of 124 meV atom -1
  • Figures 49E-G show that the products that result from the chemical reaction of LGPS and LNO (which constitute the LGPS-LNO interphase) also experience mechanically-induced metastability.
  • the lithium ions can migrate to the anode and thus form a non-local phase.
  • the local reaction dilation will be greatly reduced as the volume of the formed lithium phase will not be included in the local volume change.
  • the lithium metal phase forms locally, it contributes to a larger local volume change and, hence, a larger reaction dilation.
  • coating cathode materials in an insulator such as LNO is needed in order for constraints to lead mechanically-induced metastability on the interface of the LGPS.
  • lithium metal is soft and which leads to the difficulty of applying pressure due to the immediate short of lithium through the bulk solid electrolyte.
  • lithium metal was used as anode with a graphite layer as a protection layer, which allows high pressure applied during battery test.
  • lithium metal-LCO batteries were made at different mechanical conditions using Swagelok, aluminum pressurized cell and stainless-steel pressurized cell, as shown in Figure 57. Again, the interface reaction and decomposition reaction in the strongest constraint condition is the lowest.
  • a similar structure was applied to make a higher-voltage lithium metal battery using LCMO as cathode, where the cell was initially pressed with 6T.
  • Figure 50B depicts organic liquid electrolyte failing at nearly 5V.
  • the solid-state battery tested under isovolumetric conditions can be charged up to 9 V (Fig.50A) without evidence of a decomposition plateau.
  • a battery cycling at 5.5 V and tested under isovolumetric conditions (initially pressed with 6T) (Figure 50C) shows a stable cycling performance and high Columbic efficiency even at high cut-off voltage of 5.5 V, in contrast to the liquid battery ( Figure 50B).
  • Routine XRD data were collected in a Rigaku Miniflex 6G diffractometer working at 45 kV and 40 mA, using CuKa radiation (wavelength of 1.54056 ⁇ ). The working conditions were 2q scanning between 10–80 o, with a 0.02 o step and a scan speed of 0.24 seconds per step.
  • the LGPS+C/LGPS part of the cells were pellets which were made by pressing the powder at 1T, 3T, 6T, respectively, and put into Swagelok or the homemade pressurized cell.
  • voltage starting from the open circuit voltage to 10 V was ramped, during which the decomposition currents at each voltage were measured.
  • the CV test was conducted on a Solartron 1400 electrochemical test system between OCV to 3.2V, 7.5V, and 9.8V, respectively, with the scan rate of 0.1 mV/s. The CV scan was followed by a voltage hold for 10 hours to make sure the decomposition is fully developed, and it was scanned back to 2.5V before any other characterizations.
  • the electrochemical impedance spectroscopy (EIS) was conducted on the same machine in the range of 3 MHz to 0.1 Hz.
  • the electrode and electrolyte layers were made by a dry method which employs Polytetrafluoroethylene (PTFE) as a binder and allows to obtain films with a typical thickness of 100-200 ⁇ m.
  • PTFE Polytetrafluoroethylene
  • two different kinds of all-solid-state batteries were assembled, using Li4Ti5O12 (LTO) or lithium (Li) metal as anode.
  • the composite cathode was prepared by mixing the active materials (LiCo0.5Mn1.5O4, LiNi0.5Mn1.5O4 or LiCoO2) and Li10GeP2S12 (LGPS) powder in a weight ratio of 70:30 and 3% extra of PTFE. This mixture was then rolled into a thin film.
  • the galvanostatic battery cycling test was performed on an ArbinBT2000 work station at room temperature.
  • a Li metal foil with a diameter and thickness of 1 ⁇ 2” and 40 ⁇ m, respectively was connected to the current collector.
  • the Li foil was covered by a 5/32” diameter carbon black film with a weight ratio of carbon black and PTFE of 96:4.

Landscapes

  • Chemical & Material Sciences (AREA)
  • Electrochemistry (AREA)
  • General Chemical & Material Sciences (AREA)
  • Chemical Kinetics & Catalysis (AREA)
  • Engineering & Computer Science (AREA)
  • Manufacturing & Machinery (AREA)
  • Inorganic Chemistry (AREA)
  • Materials Engineering (AREA)
  • Physics & Mathematics (AREA)
  • Condensed Matter Physics & Semiconductors (AREA)
  • General Physics & Mathematics (AREA)
  • Secondary Cells (AREA)
  • Battery Electrode And Active Subsutance (AREA)
  • Composite Materials (AREA)

Abstract

The invention provides rechargeable batteries including a solid state electrolyte (SSE) containing an alkali metal disposed between two electrodes. The batteries are volumetrically constrained imparting increased stability under voltage cycling conditions, e.g., through microstructure mechanical constriction on the solid state electrolyte and the electrolyte-electrode interface. These batteries of the invention are advantageous as they may be all-solid-state batteries, e.g., no liquid electrolytes are necessary, and can achieve higher voltages with minimal electrolyte degradation.

Description

SOLID STATE BATTERIES
FIELD OF THE INVENTION
The invention is directed to the field of solid state batteries with alkali metal sulfide solid state electrolytes.
BACKGROUND OF THE INVENTION Solid-state lithium ion conductors, the key component to enabling all solid-state lithium ion batteries, are one of the most pursued research objectives in the battery field. The intense interest in solid-state electrolytes, and solid-state batteries more generally, stems principally from improved safety, the ability to enable new electrode materials and better low-temperature performance. Safety improvements are expected for solid-state battery cells as the currently used liquid-electrolytes are typically highly-flammable organic solvents. Replacing these electrolytes with non-flammable solids would eliminate the most problematic aspect of battery safety. Moreover, solid-electrolytes are compatible with several high energy density electrode materials that cannot be implemented with liquid-electrolyte based configurations. Solid-electrolytes also maintain better low temperature operation than liquid-electrolytes, which experience substantial ionic conductivity drops at low temperatures. Such low temperature performance is critical in the burgeoning electric-vehicles market. Of the currently studied solid-electrolytes, sulfides remain one of the highest-performance and most promising families. Sulfide glass solid-electrolytes and glass-ceramic solid-electrolytes, where crystalline phases have precipitated within a glassy matrix, have demonstrated ionic conductivities on the order of 0 ?
Figure imgf000003_0006
and above
Figure imgf000003_0002
respectively. The ceramic-sulfide electrolytes, most notably ? and ( S S) are particularly promising as they maintain
Figure imgf000003_0001
Figure imgf000003_0004
exceptionally high ionic conductivities. LGPS was one of the first solid-electrolytes to reach ionic conductivities comparable to liquid-electrolytes at
Figure imgf000003_0005
only to be displaced by LSPS, which achieved an astonishingly high ionic conductivity of Despite these promising
Figure imgf000003_0007
conductivities, the ceramic-sulfide family is plagued by a narrow stability window. That is, LGPS and LSPS both tend to reduce at voltages below approximately 1.7^? vs lithium metal or oxidize above approximately 2 This limited stability window has proven a major barrier for battery cells that need to operate in a voltage range of approximately
Figure imgf000003_0003
Thus, there is a need for improved solid state batteries incorporating solid state electrolytes with controllable structural properties and surface chemistry. SUMMARY OF THE INVENTION
We have developed rechargeable solid state batteries using solid state electrolytes with improved cycling performance. The rechargeable solid state batteries disclosed herein are advantageous as the solid state electrolytes have superior voltage stability and excellent battery cycle performance. Batteries of the invention may be stabilized against the formation of lithium dendrites and/or can operate at high current density for an extended number of cycles.
In one aspect, the invention features a rechargeable battery including a first electrode, a second electrode, and a solid state electrolyte disposed therebetween. The solid state electrolyte includes a sulfide that includes an alkali metal, such as lithium. In certain embodiments, the solid state electrolyte is under a volumetric constraint sufficient to stabilize the solid state electrolyte during electrochemical cycling. In particular embodiments, the volumetric constraint exerts a pressure of about 70 to about 1,000 MPa, e.g., about 100-250 MPa, on the solid state electrolyte, e.g., to enforce mechanical constriction on the microstructure of solid electrolyte on the order of 15 GPa. In certain embodiments, the volumetric constraint provides a voltage stability window of between 1 and 10 V, e.g., 1-8V, 5.0-8 V, or greater than 5.7 V, or even greater than 10V.
In some embodiments, the solid state electrolyte has a core shell morphology. In certain
embodiments the alkali metal is Li, Na, K, Rb, or Cs, e.g., Li. In some embodiments, the solid state electrolyte includes SiPS, GePS, SnPS, PSI, or PS. In some embodiments, the solid state electrolyte is Li10SiP2S12, Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3. In some embodiments, the first electrode is the cathode, which can include LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or LiCo0.5Mn1.5O4. In certain embodiments, the second electrode is anode and can include lithium metal, lithiated graphite, or Li4Ti5O12. In particular embodiments, the volumetric constraint provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa, e.g., about 15 GPa.
In another aspect, the invention features a rechargeable battery including a first electrode, a second electrode, and a solid state electrolyte disposed therebetween, wherein the second electrode is an anode comprising an alkali metal and graphite. In some embodiments, the battery is under a pressure of about 70-1000 MPa, e.g., about 100-250 MPa. In particular embodiments, the alkali metal and graphite form a composite. In some embodiments, the alkali metal is Li, Na, K, Rb, or Cs, e.g., Li. In some embodiments, the solid state electrolyte includes SiPS, GePS, SnPS, PSI, or PS. In certain embodiments, the solid state electrolyte is Li10SiP2S12, Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3. In particular embodiments, the first electrode is the cathode and can include LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or LiCo0.5Mn1.5O4. In some embodiments, the battery is under an external stress that provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa, e.g., about 15 GPa.
In another aspect, the invention features a rechargeable battery including a first electrode, a second electrode, and a solid state electrolyte disposed therebetween, wherein the solid state electrolyte may include a sulfide including an alkali metal; and the battery is under isovolumetric constraint. In some embodiments, the isovolumetric constraint is provided by compressing the solid state electrolyte under a pressure of about 3-1000 MPa, e.g., about 100-250 MPa. In certain embodiments, the alkali metal is Li, Na, K, Rb, or Cs, e.g., Li. In some embodiments, the solid state electrolyte includes SiPS, GePS, SnPS, PSI, or PS. In certain embodiments, the solid state electrolyte is Li10SiP2S12, Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3. In particular embodiments, the first electrode is the cathode and can include LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or LiCo0.5Mn1.5O4. In some embodiments, the isovolumetric constraint provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa, e.g., about 15 GPa. In another aspect, the invention features a rechargeable battery having a first electrode, a second electrode, and a solid state electrolyte disposed therebetween. The solid state electrolyte includes a sulfide that includes an alkali metal, and optionally has a core-shell morphology. The first electrode includes an interfacially stabilizing coating material. In certain embodiments, the first and second electrodes independently include an interfacially stabilizing coating material. In certain embodiments, one of the first and second electrodes includes a lithium-graphite composite.
In some embodiments, the first electrode comprises a material as described herein, e.g., in Table 1. In some embodiments, the coating material of the first electrode is a coating material as described herein, e.g., LiNbO3, AlF3, MgF2, Al2O3, SiO2, graphite, or in Table 2. In certain embodiments, the alkali metal is Li, Na, K, Rb, or Cs, e.g., Li. In some embodiments the solid state electrolyte includes SiPS, GePS, SnPS, PSI, or PS. In certain embodiments, the solid state electrolyte is Li10SiP2S12, Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3. In some embodiments, the first electrode is the cathode and can include LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or
LiCo0.5Mn1.5O4. In some embodiments, the battery is under an external stress that provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa, e.g., about 15 GPa. In certain embodiments, the battery is under a pressure of about 70-1000 MPa, e.g., about 100-250 MPa.
In another aspect, the invention features a method of storing energy by applying a voltage across the first and second electrodes and charging the rechargeable battery of the invention. In another aspect, the invention provides a method of providing energy by connecting a load to the first and second electrodes and allowing the rechargeable battery of the invention to discharge.
BRIEF DESCRIPTION OF THE DRAWINGS
Figures 1A-1B: Cyclic Voltammetry (CV) tests of LGPS in liquid (A) and solid (B) states at different pressures. LGPS/C thin film with the ratio of 90:10 was tested in the liquid electrolyte (black curve in (A)). The CV tests were also conducted by replacing liquid electrolyte with LGPS pellets, which is all- solid-state CV, at different pressures. The decomposition intensity is decreased significantly with increasing applied pressure. At a reasonably low pressure of 6 T (420 MPa), there is already no notable decomposition peaks before 5.7 V (purple curve), which indicates applying external pressure or volume constriction on the battery cell level can widen the electrochemical window of the solid- state electrolyte. Figures 2A-2B: Capacity (A) and cycling performance (B) of LiCoO2 (LCO)- Li4Ti5O12 (LTO) all-solid- state full battery. As the chemical potential of LTO is 1.5 V (vs. Li), the working plateau in cathode side is higher than 4 V (vs. Li).
Figures 3A-3B: Capacity (A) and cycling performance (B) of LiNi0.5Mn1.5O4 (LNMO)-LTO all-solid- state full battery. As the chemical potential of LTO is 1.5 V (vs. Li), the working plateau in cathode side is higher than 4.7 V (vs. Li).
Figure 4: High voltage cathode candidates for 6V and greater all solid state Li-ion battery technology. The legend labels are: F are fluorides, O are oxides, P,O are phosphates, and S,O: sulfates. The complete list of these high voltage fluorides, oxides, phosphates, and sulfates is provided in Table 1. Commercial LiCoO2 (LCO) and LMNO are labeled as stars.
Figs 5A-5B: (A) Illustration of the impact of strain on LGPS decomposition, where ?? is the fraction of LGPS that has decomposed. The lower dashed line represents the Gibbs energy (??????) of a binary combination of pristine LGPS and an arbitrary set of decay products (?) when negligible pressure is applied (isobaric decay with
Figure imgf000006_0010
The solid line shows the Gibbs when a mechanical constraint is applied to the LGPS. Since LGPS tends to expand upon decomposition, the strain Gibbs (???????) increases when such a mechanical constraint is applied. At some fracture point, denoted ??, the Gibbs energy of the system exceeds the energy needed to fracture the mechanical constraints (the upper dashed line). The highlighted path is the suggested ground state for a mechanically constrained LGPS system. The region
Figure imgf000006_0008
is metastable if Schematic representation of work
Figure imgf000006_0009
differentials in the cases of“fluid" and "solid" like systems. For the top,“fluid-like”, system, the system undergoes an internal volume expansion due to decomposition rather than an applied stress (“stress- free” strain). The bottom system represents the elastic deformation away from an arbitrary reference state. Figure 6: Stability windows for LGPS and
Figure imgf000006_0002
in the mean field limit.
Figure imgf000006_0004
Figure imgf000006_0001
indicates how rigid the constraining mechanism is. The limits
Figure imgf000006_0003
¥^represent the isovolumetric and isobaric limits. In the isobaric case, the intrinsic material stability 1 V) is recovered. Figures 7A-7B: (A) Illustration of the nucleated decay mechanism. A pristine LGPS particle of radius R? undergoes a decay within a region of radius R? at its center. The decomposed region’s radius in the absence of stress is now R?, which must be squeezed into the void of R?. The final result is a nucleated particle (iv) where the strain is non-zero. (B)
Figure imgf000006_0005
in units of KV for both the hydrostatic/mean field and nucleated models. For typical Poisson ratios, it is seen that the strain term is comparable to or better than an ideal core-shell model
Figure imgf000006_0006
Figures 8A-8E: Voltage (j), lithium chemical potential ( ) and Fermi level
Figure imgf000006_0007
distributions in various battery configurations. (A) Conventional battery design. (B) Conventional battery with hybrid solid-electrolyte/active material cathode. c? gives the interface voltage that forms between the active material and the solid-electrolyte because of the different lithium ion chemical potentials. (C) Illustration of previous speculation of how insulating layers could lead to variable lithium metal chemical potentials within the cell. (D) Expectation of how the voltage from part (C) would relax given the effective electronic conduction that occurs due to lithium hole migration. (E) The result of part (D) once the applied voltage exceeds the intrinsic stability window of the solid-electrolyte. Local lithium is seen to form within the insulated region with an interface voltage (c?) equal to the applied voltage. Figures 9A-9D: Comparison between microstructures and chemical composition of LGPS and ultra- LGPS particles. (A, C) Typical TEM bright-field images of LGPS and ultra-LGPS particles respectively, showing a distinct surface layer for ultra-LGPS particle. (B, D) Statistically analyzed STEM EDS linescans performed on various LGPS and ultra-LGPS particles with different sizes, showing a uniform distribution of sulfur concentration from surface to bulk for LGPS particles, but a decreased sulfur concentration in surface layer for ultra-LGPS.
Figure 10: STEM EDS linescans across individual LGPS particles with different particle sizes ranging from 100nm to 3µm, showing that the sulfur concentration variation from surface to the bulk has no regular pattern.
Figure 11: STEM EDS linescans across individual LGPS particles sonicated in dimethyl carbonate (DMC) for 70h with different particle sizes ranging from 60nm to 4µm, showing that sulfur concentration is obviously smaller at surface region compared to that in the bulk.
Figure 12: STEM EDS linescans across individual LGPS particles sonicated in diethyl carbonate (DEC) for 70h with different particle sizes ranging from 120nm to 4µm, showing that sulfur concentration is obviously smaller at surface region compared to that in the bulk.
Figure 13: Quantitative STEM EDX analyses of LGPS particles before and after ultrasonic preparation show that surface/bulk ratio of S is obviously lower after sonication in organic electrolytes (DEC and DMC).
Figure 14: STEM EDS linescans across individual LGPS particles soaked in DMC for 70h without sonication with different particle sizes ranging from 160nm to 3µm, showing that the sulfur concentration variation from surface to the bulk has no regular pattern.
Figures 15A-15H: Comparison between electrochemical performances of LGPS and ultra-LGPS particles, and LIBs made from LGPS and ultra-LGPS particles. (A, B) Cyclic voltammograms(CV) of Li/LGPS/LGPS+C/Ta and Li/ultra-LGPS/ultra-LGPS/Ta cells respectively, with a lithium reference electrode at a scan rate of 0.1mVs-1 and a scan range of 0.5 to 5 V. (C, D) Sensitive electrochemical impedance spectra (EIS) for LGPS and ultra-LGPS cells in panel (A,B) before and after CV tests. (E, F) Charge-discharge profiles of LGPS-LIB (LTO+LGPS+C/Glass fiber separator/Li) and ultra-LGPS- LIB (LTO+ultra-LGPS+C/Glass fiber separator/Li) cycled at 0.5C current rate in the voltage range of 1.0 - 2.2 V. (G, H) Cyclic capacity curves of LGPS LIB and ultra-LGPS-LIB.
Figures 16A-16B: Cycling performance of (A) LGPS-ASSLIB (LTO+LGPS+C as cathode, LGPS as solid electrolyte, and Li as anode) and (B) ultra-LGPS-ASSLIB (LTO+ultra-LGPS+C as cathode, ultra- LGPS as solid electrolyte, and Li as anode) at low current rate (0.02C). Figures 17A-17B: Cycling performance of (A) LGPS-ASSLIB (LTO+LGPS+C as cathode, LGPS as solid electrolyte, and Li as anode) and (B) ultra-LGPS-ASSLIB (LTO+ultra-LGPS+C as cathode, ultra- LGPS as solid electrolyte, and Li as anode) at medium current rate (0.1C).
Figure 18A-18B: Cycling performance of (A) LGPS-ASSLIB (LTO+LGPS+C as cathode, LGPS as solid electrolyte, and Li as anode) and (B) ultra-LGPS-ASSLIB (LTO+ultra-LGPS+C as cathode, ultra- LGPS as solid electrolyte, and Li as anode) at high current rate (0.8C).
Figures 19A-19G: Microstructural and compositional (S)TEM studies of LTO/LGPS interfaces after cycling in LGPS ASSLIB. (A) FIB sample prepared from LGPS ASSLIB after 1 charge- discharge cycle, in which the cathode layer (LTO+LGPS+C) and SE layer (LGPS) are included. (B) TEM BF images of LTO/LGPS primary interface, showing a transit layer with multiple dark particles. (C) HRTEM image of LTO particle and its corresponding FFT pattern. (D) STEM DF image of LTO/LGPS primary interface shows super bright particles within the transit layer, indicating the accumulation of heavy elements. (E) STEM EELS linescans performed across the primary interface, indicating that the bright particles within the transit layer are sulfur-rich. (F) STEM DF image of LTO/LGPS secondary interface, in which a higher density of bright particles with similar morphology show up again. (G) STEM EELS linescans performed across the secondary interface, indicating that the bright particles are sulfur-rich.
Figure 20: TEM bright-field images and STEM dark-field image of primary LTO/LGPS interface (interface between cathode and LGPS solid electrolyte layer) of LGPS-ASSLIB (LTO+LGPS+C as cathode, LGPS as solid electrolyte, and Li as anode), showing an obvious transit layer between the cathode and solid electrolyte layer.
Figures 21A-21B: (A) STEM dark-field image of and (B) EELS linescan on primary LTO/LGPS interface (interface between cathode and LGPS solid electrolyte layer) of LGPS-ASSLIB
(LTO+LGPS+C as cathode, LGPS as solid electrolyte, and Li as anode), showing that LiK and GeM4,5 peaks exist for regions both inside and outside bright particles within the transit layer.
Figures 22A-22B: (A) STEM dark-field image of and (B) EELS linescan on primary LTO/LGPS interface (interface between cathode and LGPS solid electrolyte layer) of LGPS-ASSLIB
(LTO+LGPS+C as cathode, LGPS as solid electrolyte, and Li as anode), showing that SL1 peak intensity is stronger on those S-rich bright-contrast particles within the transit layer.
Figures 23A-23F: Microstructural and compositional (S)TEM studies of LTO/ultra-LGPS interfaces after cycling in ultra-LGPS ASSLIB. (A) TEM BF image of LTO/ultra-LGPS primary interface, showing a smooth interface with no dark particles that exist in Figure 6B. (B) STEM EELS linescan spectra corresponding to the dashed arrow in Figure 23A. (C) STEM DF image of LTO/ultra-LGPS secondary interface. (D) STEM EDS linescans show a continuously decreasing atomic percentage of sulfur from inner ultra-LGPS particle to secondary LTO/ultra-LGPS interface, and finally into LTO+C composite region. (E) STEM EDS mapping shows that the large particle in Figure 22C is LGPS particle. (F) STEM EDS quantitative analyses show that the atomic percentage of sulfur inside ultra-LGPS particle is as high as ~38%, while that of secondary LTO/ultra-LGPS interface is as low as 8%. Figure 24A-24B: Additional (A) STEM dark filed images and (B) STEM EDX linescans showing a much lower S concentration at the secondary LTO/ultra-LGPS interface than inner ultra-LGPS particle region.
Figure 25A-25C: (A) The number of hulls required to evaluate the stability of the 67k materials considered if the evaluation schema is material iteration (left columns) or elemental set iteration (right columns). (B) An illustration of the pseudo-binary approach to interfacial stability between LSPS and an arbitrary material A. represents the materials-level decomposition energy that exists even in
Figure imgf000009_0011
the absence of the interface, whereas
Figure imgf000009_0012
represents the added instability due to the presence of the interface. The most kinetically driven reaction occurs when ?
Figure imgf000009_0001
and
Figure imgf000009_0014
are the decomposed coating material and LSPS in the absence of an interface
Figure imgf000009_0013
(C) Correlation of elemental fraction with the added chemical interfacial instability . Negative values are those
Figure imgf000009_0002
atomic species such that increasing the concentration decreases and improves interfacial
Figure imgf000009_0003
stability. Conversely, positive values are those atomic species that tend to increase G?
???? and worsen interfacial stability. Elements that are only present in less than 50 crystal structures are grayed out due to lack of high-volume data.
Figures 26A-26C: (A-C) Correlation of elemental species fraction with the added electrochemical interfacial instability
Figure imgf000009_0004
( ???? ) at 0, 2 and 4 V, respectively. Negative values are those species such that increasing concentration decreases
Figure imgf000009_0005
and improves interfacial stability. Conversely, positive values are those species that tend to increase and worsen interfacial stability. Elements that are
Figure imgf000009_0006
only present in less than 50 crystal structures are grayed out due to lack of high-volume data.
Figures 27A-27D: (A) Hull energy vs voltage relative to lithium metal for LSPS. Darker Gray [Mid- Gray] shading highlights where the decomposition is oxidative [reductive]. Light gray shading represents the region where LSPS decays to without consuming or producing lithium (e.g. lithium neutral). The oxidation [reduction] region is characterized by a hull energy that increases [decreases] with increasing voltage. (B) and (C) Hull energies at the boundary voltages for the anode and cathode ranges, respectively, in terms of anionic species (e.g., oxygen containing compounds vs sulfur containing compounds, etc.). Data points below [above] the neutral decay line are net oxidative
[reductive] in the anode/cathode ranges. Those compounds on the neutral decay line are decaying without reacting with the lithium ion reservoir. (D) Average hull energy for material-level
electrochemical decompositions versus voltage.
Figures 28A-28C: Comparison of average LSPS interfacial stability of compounds sorted by anionic species. (A) The average total maximum kinetic driving energy and the contribution due to
Figure imgf000009_0009
the interface
Figure imgf000009_0007
(G? ? ?) for chemical reactions between LSPS and each of the considered anionic classes. (B) The total electrochemical instability
Figure imgf000009_0010
of each anionic class at a given voltage. (C) The average contribution of the interface to the electrochemical instability of each
Figure imgf000009_0008
anionic class at a given voltage.
Figures 29A-29B: Functionally stable results for compounds sorted by anionic species. (A) and (B) The total number (line) and percentage (bar) of each anionic class that was determined to be functionally stable. The bottom bar represented the percentage of materials that are functionally stable and the top bar represents the percentage of materials that are potentially functionally stable depending on the reversibility of lithiation/delithiation.
Figures 30A-30F: (A-D) Comparison of XRD patterns to show structural decay of LCO, SnO2, LTO and SiO2 at the solid-electrolyte material interface (with no applied voltage). In (A) ,^, , , ,^ stand for LCO(PDF# 44-0145), LSPS(ICSD#252037), SiO2(PDF# 48-0476), Li3PO4(PDF# 45-0747), Cubic Co4S3(PDF# 02-1338), Monoclinic Co4S3(PDF# 02-1458) respectively. In (B), ,^, , ,^ stand for SnO2(PDF#41-1445), LSPS(ICSD#252037), SiO2(PDF# 34-1382), P2S5 (PDF# 50-0813), and Li2S(PDF# 23-0369) respectively. In (C), ,^, , stand for LTO(PDF#49-0207), LSPS(ICSD#252037) and Li1.95Ti2.05S4 (PDF# 40-0878) respectively. In (D), ,^ stand for SiO2(PDF#27-0605) and
LSPS(ICSD#252037) respectively. The shaded regions in (A-D) highlight where significant phase change happened after heating to 500 °C. The interfacial chemical compatibility decreases from (A) to (D), corresponding well with the predicted interfacial decay energies of 200, 97, 75, and 0 meV/atom for LCO, SnO2, LTO and SiO2, respectively. (E, F) CV results for Li?S and SnO?. The shaded regions predict if the curve in that region will be dominantly oxidation, reduction, neutral. Figures 31A-31E: Comparison of XRD patterns for each individual phase: (A) LiCoO2, (B) LSPS, (C) Li4Ti5O12, (D) SnO2 and (E) SiO2, at room temperature and 500°C. No significant change between room temperature and 500°C can be observed for each phase.
Figures 32A-32D: Comparison of XRD patterns for mixture powders: (A) LiCoO2+LSPS, (B)
SnO2+LSPS, (C) Li4Ti5O12+LSPS, and (D) SiO2+LSPS) at various temperatures (room temperature, 300°C, 400°C and 500°C). The onset reaction temperature is observed to be 500°C, 400°C and 500°C for LiCoO2+LSPS, SnO2+LSPS and Li4Ti5O12+LSPS, respectively. No reaction is observed to happen for SiO2+LSPS up to 500°C.
Figures.33A-33F (A, B, C) XRD of different powder mixtures before and after heat treatment at 500℃ for 36 hours ((A) Li + LGPS; (B) Graphite + LGPS; (C) Lithiated graphite + LGPS). The symbols and corresponding phases are:
Figure imgf000010_0001
) The structure of Li/Graphite anode in LGPS based all-solid-state battery; (E) SEM image of the cross section of Li/Graphite anode; (F) FIB-SEM of the interface of Li and Graphite.
Figures.34A-34E (A) The comparison of cyclic performance between Li/G-LGPS-G/Li and Li-LGPS-Li symmetric batteries; (B) The SEM images of symmetric batteries after cycling. Li/G-LGPS-G/Li symmetric battery after 300 hours’ cycling (B1,2) and Li-LGPS-Li symmetric battery after 10 hours’ cycling (B3,4); (C) The rate performance of Li/G-LGPS-G/Li symmetric batteries under different pressures. (D) The SEM images of Li/G-LGPS-G/Li symmetric batteries under different pressures after rate tests. (E) The ultra-high rate performance up to 10 mA/cm2 of Li/G-LGPS-G/Li symmetric batteries. The pressure applied in (E) is 250 MPa. Insets are the cycling profiles plotted in the range of -0.3V to 0.3V, showing that there is no obvious change of overpotential after high rate cycling. More voltage profile enlargements are shown in supplementary information Figure 42. Figures.35A-35D (A) The comparison of initial charge/ discharge curves, (B) the initial Coulombic efficiencies and (C) the open circuit voltages after 1h rest, among different capacity ratios of Li to Graphite in Li/G-LGPS-LCO (LiNbO3 coated) system. The Li/G capacity ratio of 0, 0.5, 0.8, 1.5, 2.5 and 4 can be translated into Li/G thickness ratio of around 0, 0.3, 0.4, 0.8, 1.3, and 2.1 respectively. Without specific explanation, the Li/graphite thickness ratio is 1.0-1.3 by default in this work. (D) Cyclic performance of Li/G-LGPS-LCO (LiNbO3 coated) battery.
Figures 36A-36B. (A) Voltage profiles of LGPS decomposition at different effective modules (Keff). (B) Reduction reaction pathways corresponding to different Keff and the products in different phase equilibria within each voltage range. All decomposition products here are the ground state phases within each voltage range.
Figures 37A-37F. XPS measurement of Ge and P for anode-LGPS-anode symmetric batteries with the X-ray beam focused on (A) the center part LGPS away from the interface to Li/G and (B) the interface between Li/G and LGPS in Li/G-LGPS-G/Li cell under 100 MPa after 12 hours cycle at 0.25 mA cm-2; (C) the interface between Li and LGPS in Li-LGPS-Li symmetric battery under 100 MPa after 10 hours cycles at 0.25 mA cm-2 (failed); (D) The Li/G-LGPS interface after rate test at 2 mA cm-2 under 100 MPa and (E) 10 mA cm-2 under 250 MPa; (F) The Li/G-LGPS interface at 2 mA cm-2 under 3 MPa.
Figures 38. XRDs of graphite and the mixture of Li and graphite after heating under 500℃ for 36 h. Figures 39A-39C. SEM images of (A) graphite particles; the surface (B) and cross section (C) of graphite film after applying high pressure.
Figures 40. Cyclic performance of Li/G-LGPS-G/Li symmetric battery with relatively smaller overpotential.
Figures 41A-44B. Comparison of SEM images of Li/G anode before (A) and after (B) long-term cycling in Figure 34(A).
Figures 42A-42C. (A) Rate test of Li/G-LGPS-G/Li symmetric battery. When the pre-cycling time is reduced to 5 cycles at 0.25 mA cm-2, the battery“fails” at 6 mA cm-2 or 7 mA cm-2, however, when the current density is set back to 0.25 mA cm-2, it always comes back normal without significant overpotential increase. (B) Enlarged Figure 34(E2), battery cycled at 10 mA cm-2 plotted in a smaller voltage scale (B1) or time scale (B2). (C) SEM images of Li/Graphite composite after testing showing in B with different area and magnification. No lithium dendrite was observed. A clear 3D structure showing this is in Figure 42(C2).
Figures 43A-43B. (A) cycling profiles of LCO-LGPS-Li/G batteries in Figure 35D. (B) Cyclic performance based on Li anode. Both batteries were tested at current density of 0.1 C at 25℃.
Figures 44A-44B. Bader charge analysis from DFT simulations. (A) Phosphorus element in all the P- related compounds from the decomposition product list; (B) Ge element in all the Ge-related compounds from the decomposition product list. Figures 45A-45D. (A) Comparison of CV curves of Li/G-LGPS-LGPS/C battery tested under 3 and 100 MPa; (B,C) comparison of impedance change before and after these two CV tests; (D) Model used in impedance fitting. Rbulk stands for the ionic diffusion resistance and Rct represents the charge transfer resistance. All EIS data are fitted with Z-view.
Figures 46A-46G. (A) A CV test of Swagelok battery after they are pressed with 1T, 3T, 6T and pressurized cell initially pressed with 6T.10 % carbon is added in the cathode. The voltage range is set from open circuit to 9.8 V. (B) The CV scans in (A) plotted in a magnified voltage and current ranges. (C) In-situ impedance tests during CV scans for batteries shown in (A). (D) Synchrotron XRD of pressurized cells after no electrochemical process (black), CV scan to 3.2V, 7.5V and 9.8V. All CVs were followed by a voltage holding at the same high cutoff voltages for 10 hours and then discharged back to 2.5V. Green line: Synchrotron XRD of LGPS tested in liquid electrolyte after CV scan to 3.2V and held for 10 hours. (E) Synchrotron XRD peak of different batteries at 2q=18.5°, showing the broadening of XRD peak after high-voltage CV scan and hold. (F) Strain versus size broadening analysis for LGPS after high voltage hold. Dots are the broadening of different peaks in 7.5V SXRD measurement, with the corresponding XRD peaks shown in Figure 52. The angle dependences of size and strain broadenings are represented by dashed lines. (G) XAS measurement of S (g1) and P (g2) after high voltage CV scan and hold. (g3) The simulation of P XAS peak shift after straining in the c-direction.
Figures 47A-47D. (A) LGPS decomposition energy (a1), ground state pressure (a2), and ground state capacity versus voltage at different effective modules (Keff). (B) Decomposition reaction pathways at different Keff and the products induced by different phase equilibriums in different voltage ranges. (C,D) XPS measurement of S (c) and P (d) element for pristine LGPS (c1, d1), battery after 3.2 V CV scan in liquid electrolyte (c2, d2), pressurized cell after 3.2 V CV scan (c3, d3) and pressurized cell after 9.8 V CV scan (c4, d4). Each CV scan is followed by a 10 hour hold at the high cutoff voltage. Figures 48A-48E. Galvanostatic charge and discharge voltage curves for all-solid-state batteries using: (A1) LCO, (A2) LNMO and (A3) LCMO as cathode material versus LTO. The cyclability of the batteries is represented in (B1), (B2) and (B3) for LCO, LNMO and LCMO, respectively. Here, LCO and LNMO are charged and discharged at 0.3C, whereas LCMO is charged at 0.3 C and discharged at 0.1 C. All batteries are tested at room temperature, in the pressurize cell initially pressed with 6T and activate materials are coated with LiNbO3, as shown in Figure 54. (C,D) XPS measurement of LCO, LNMO, LCMO-LGPS before and after 5 cycles. (E) XAS measurement of LCO, LNMO, LCMO- LGPS before (E1) and after (E2) 5 cycles for element S.
Figures 49A-49G. (A-D) Pseudo phase simulations of the interface between LGPS and (A) LNO, (B) LCO, (C) LCMO, (D) LNMO. Plots depict the reaction energy of the interface versus the atomic fraction of the non-LGPS phase consumed. The value of the atomic fraction that has the most severe decomposition energy is defined to be ??. (E-G) Mechanically-induced metastability plots for the LGPS-LNO interphase (the set of products that result from the decomposition in Figure 49A). (E) Energy over hull of the interphase show significant response to mechanical constriction. (D) and (E) Show analogous behavior to the pressure and capacity responses to pressure that were observed for bulk phase LGPS (Figures 47A-47D).
Figures 50A-50C. (A) Galvanostatic charge and discharge profiles for all-solid-state batteries using LCO and LCMO as cathode and graphite coated lithium metal as anode, with cut-off voltage from 2.6- 4.5 V(LCO) and 2.6- (6-9) V (LCMO).The batteries are charged at 0.3C and discharged at 0.1C. Cycling performance of LCMO lithium metal battery using (B) 1M LiPF6 in EC/DMC and (C) constrained LGPS as electrolyte, with cut-off voltage from 2.5-5.5V with charge rate of 0.3C and discharge rate of 0.1 C.
Figure 51. Pellet thickness change in response of force applied. The original thickness of pellet is 756 µm, the weight of the pellet is 0.14 g, the area of the pellet is 1.266 cm2, the compressed thickness of the pellet is 250 µm. the calculated density is 2.1 g/cm3, which is close to the theoretical density of LGPS of 2 g/cm3.
Figures 52A-52F. (A)-(F) Synchrotron XRD peaks of batteries at different 2q angles, showing the broadening of XRD peak after high-voltage CV scan and hold. The pressurized cell after 3.2V CV scan and hold doesn’t show XRD broadening. Figure 53. (top) Illustration of decomposition front propagation. Decomposed phases are marked with ?…?. Such propagation is seen to require tangential ionic conduction. (bottom) Energy landscape for reaction coordinates. The final result is a shift in Gibbs energy by D?, which is positive or negative based on equation 2. Even when D? is negative (reaction is thermodynamically favorable), the presence of a sufficient overpotential due to tangential currents can significantly reduce the front’s propagation rate.
Figure 54. STEM image and EDS maps of LiNbO3 coated LCO.
Figure 55. Rate testing of LCO-LTO battery using LGPS thin film as electrolyte, battery was tested at 0.3 C-2.5 C.
Figure 56. XAS measurement of LCO, LNMO, LCMO-LGPS before (represented as p) and after (represented as 5c) 5 cycles for element P.
Figures 57A-57B (A) Charge and (B) discharge profiles of LCO all-solid-state batteries using LGPS as electrolyte tested with Swagelok, Al pressurized cell, and Stainless steel (SS) pressurized cell with voltage cut-off between 3V-4.15V. Swagelok applied almost no pressure; Al cell is soft compared with Stainless steel and which applied low constrain while stainless steel applied the strongest constant constrain during battery test.
Figures 58A-58B. Comparison of CV current density of LGPS+Cathode and LGPS+C. CV
measurement of LGPS+LCO (30+70) (A) and LGPS+LCMO (30+70) (B) in pressurized cells and CV measurement of LGPS+C (90+10) in pressurized cells.
Figures 59A-59D LCMO/LGPS/Li all-solid-state batteries assembled with (A) bare lithium metal, (B) graphite and (C) graphite coated Li as anode. (D) Cycling performance of LCMO solid battery using different anodes. At first cycle, all the three sample could be charged to around 120 mAh/g, while apparently Li/graphite shows the highest discharging capacity at about 83 mAh/g. It is clear to see that both of Li and Graphite anode suffer from quick fading within the first 5 cycles and after 20 cycles, both of their capacities dropped below 20 mAh/g. In comparison, the capacity of Li/Graphite anode maintains.
DETAILED DESCRIPTION OF THE INVENTION
The invention provides rechargeable batteries including a solid state electrolyte (SSE) containing an alkali metal and a sulfide disposed between two electrodes. The solid state electrolytes may have a core-shell morphology, imparting increased stability under voltage cycling conditions. These batteries of the invention are advantageous as they may be all-solid-state batteries, e.g., no liquid electrolytes are necessary, and can achieve higher voltages with minimal electrolyte degradation. Core-shell morphologies in which a core of ceramic-sulfide solid-electrolyte is encased in a rigid amorphous shell have been shown to improve the stability window. The mechanism behind this stabilization is believed to be tied to the tendency of ceramic-sulfides to expand during decay by up to more than 20%. Applying a volume constraining mechanism, this expansion is resisted which in turn inhibits decay. We have generalized this theory and provide experimental evidence using post- synthesis creation of a core-shell morphology of LGPS to show improved stability. Based on the decay morphology, the magnitude of stabilization will vary. A mean-field solution to a generalized strain model is shown to be the lower limit on the strain induced stability. The second decay morphology explored, nucleated decay, is shown to provide a greater capability for stabilization. Moreover, experimental evidence suggests the decay is in fact the later (nucleated) morphology, leading to significant potential for ceramic-sulfide full cell batteries. Further developments of the theory underpinning the enhanced stability and performance of core- shell electrolytes have revealed that the strain stabilization mechanism is not limited to the materials level but can also be applied on the battery cell level through external stress or volume constriction. The strain provided by the core-shell structure stabilizes the solid electrolyte through a local energy barrier, which prevents the global decomposition from happening. Such stabilization effect provided by local energy barrier can also be created by applying an external stress or volume constriction from the battery cell, where up to 5.7 V voltage stability window on LGPS can be obtained as shown in Figures 1A-1B. Higher voltage stability window beyond 5.7 V can be expected with higher pressure or volume constriction in the battery cell design based on this technology.
In solid state batteries, lithium dendrites form when the applied current density is higher than a critical value. The critical current density is often reported as 1-2 mA cm-2 at an external pressure of around 10 MPa. In the present invention, a decomposition pathway of the solid state electrolyte, e.g., LGPS, at the anode interface is modified by mechanical constriction, and the growth of lithium dendrite is inhibited, leading to excellent rate and cycling performances. No short-circuit or lithium dendrite formation is observed after the batteries are cycled at a current density up to 10 mA cm-2. Solid State Electrolytes
A rechargeable battery of the invention includes a solid electrolyte material and an alkali metal atom incorporated within the solid electrolyte material. In particular, solid state electrolytes for use in batteries of the invention may have a core-shell morphology, with the core and shell typically having different atomic compositions. Suitable solid state electrolyte materials include sulfide solid electrolytes, e.g., SixPySz, e.g., SiP2S12 such as Li10SiP2S12, or b/g-PS4. Other solid state electrolytes include, but are not limited to, germanium solid electrolytes, e.g., GeaPbSc, e.g., GeP2S12 such as Li10GeP2S12, tin solid electrolytes, e.g., SndPeSf, e.g., SnP2S12, iodine solid electrolytes, e.g., P2S8I crystals, glass electrolytes, e.g., alkali metal-sulfide-P2S5 electrolytes or alkali metal-sulfide-P2S5- alkali metal-halide electrolytes, or glass- ceramic electrolytes, e.g., alkali metal-PgSh-i electrolytes. Another material includes
Li9.54Si1.74P1.44S11.7Cl0.3. Other solid state electrolyte materials are known in the art. The solid state electrolyte material may be in various forms, such as a powder, particle, or solid sheet. An exemplary form is a powder. Alkali metals useful for the solid state electrolytes for use in batteries of the invention include Li, Na, K, Rb, and Cs, e.g., Li. Examples of Li-containing solid electrolytes include, but are not limited to, lithium glasses, e.g., xLi2S·(1-x)P2S5, e.g., 2Li2S·P2S5, and xLi2S·(1-x)P2S5–LiI, and lithium glass- ceramic electrolytes, e.g., Li7P3S11-z. Electrode Materials
Electrode materials can be chosen to have optimum properties for ion transport. Electrodes for use in a solid state electrolyte battery include metals, e.g., transition metals, e.g., Au, alkali metals, e.g., Li, or crystalline compounds, e.g., lithium titanate such as Li4Ti5O12 (LTO). An anode may also include a graphite composite, e.g., lithiated graphite. Other materials for use as electrodes in solid state electrolyte batteries are known in the art. The electrodes may be a solid piece of the material, or alternatively, may be deposited on an appropriate substrate, e.g., a fluoropolymer or carbon. For example, liquefied polytetrafluoroethylene (PTFE) has been used as the binder when making solutions of electrode materials for deposition onto a substrate. Other binders are known in the art. The electrode material can be used without any additives. Alternatively, the electrode material may have additives to enhance its physical and/or ion conducting properties. For example, the electrode materials may have an additive that modifies the surface area exposed to the solid electrolyte, such as carbon. Other additives are known in the art.
High voltage cathodes of 4 volt LiCoO2 (LCO, shown in Figures 2A-2B) and 4.8V LiNi0.5Mn1.5O4 (LNMO, shown in Figures 3A-3B) are demonstrated to run well in all-solid-state batteries of the invention. Higher voltage cathodes, such as the 5.0V Li2CoPO4F, 5.2V LiNiPO4, 5.3V Li2Ni(PO4)F, and 6V LiMnF4 and LiFeF4 may also be used as electrode materials in all-solid-state batteries of the invention. Voltage stability windows beyond 5.7 V, e.g., up to 8 or 10 V or even higher, may be achieved. Another cathode is LiCo0.5Mn1.5O4 (LCMO). Exemplary cathode materials are listed in Table 1, with the calculated stability of the electrodes in Table 1 shown in Figure 4.
Table 1: High voltage (greater than 6 V) electrode candidates with individual Materials Project Identifiers.
Figure imgf000017_0001
Figure imgf000018_0001
181. Li4Mn4F16: mp-776813 182. Li2Fe2F8:
mp-776881 183. Li4Fe4F16:
mp-777008 184. Li4Mn2F12:
mp-777332 185. Li6Fe2F12:
mp-777459 186. Li4Fe4F16:
mp-777875 187. Li4Fe2F10:
mp-778345 188. Li4Fe4F16:
mp-778347 189. Li4Mn2F12:
mp-778394 190. Li4Fe4F16:
mp-778510 191. Li4Mn4F16:
mp-778687 192. Li4Ge2F12:
mp-7791 193. Li4Mn4F16:
mp-780919
Electrode Coatings
In some cases, the electrode materials may further include a coating on their surface to act as an interfacial layer between the base electrode material and the solid state electrolyte. In particular, the coatings are configured to improve the interface stability between the electrode, e.g., the cathode, and the solid electrolyte for superior cycling performance. For example, coating materials for electrodes of the invention include, but are not limited to graphite, LiNbO3, AlF3, MgF2, Al2O3, and SiO2, in particular LiNbO3 or graphite.
Based on a new high-throughput analysis schema to efficiently implement computational search to very large datasets, a library of different materials was searched to find those coating materials that can best stabilize the interface between sulfide solid-electrolytes and typical electrode materials, using as an example to predict over 1,000 coating materials for cathodes and over 2,000
Figure imgf000020_0001
coating materials for anodes with both the required chemical and electrochemical stability. These are generally applicable for LGPS. Table 2 provides the predicted effective coating materials.
Figure imgf000021_0001
Figure imgf000022_0001
Figure imgf000023_0001
Figure imgf000024_0001
Figure imgf000025_0001
Figure imgf000026_0001
Zr1Ru1: mp-214 Zr1Zn1: mp-570276 Zr1Zn1Cu2: mp-11366 Zr1Zn1Ni4: mp-11533 Zr1Zn1Rh2: mp-977582 Zr2Be2Si2: mp-10200 Zr2Si2: mp-11322 Zr2Ti2As2: mp-30147 Zr2V2Si2: mp-5541 Zr3Cu4Ge2: mp-15985 Zr3Si2Cu4: mp-7930 Zr4Co4P4: mp-8418 Zr4Mn4P4: mp-20147 Zr4Si4: mp-893
Zr4Si4Pt4: mp-972187 Zr4V4P4: mp-22302
Figure imgf000028_0001
Figure imgf000029_0001
Figure imgf000030_0001
Figure imgf000031_0001
Figure imgf000032_0001
Figure imgf000033_0001
Figure imgf000034_0001
Figure imgf000035_0001
Figure imgf000036_0001
Figure imgf000037_0001
Figure imgf000038_0001
Figure imgf000039_0001
Figure imgf000040_0001
Figure imgf000041_0001
Figure imgf000042_0001
Figure imgf000043_0001
Figure imgf000044_0001
Figure imgf000045_0001
Figure imgf000046_0001
External Stress
Strain stabilization mechanism for enhancing electrolyte stability is not limited to the materials level but can also be applied on the battery cell level through external stress or volume constriction. In certain embodiments, the external stress is a volumetric constraint applied to all or a portion, e.g., the solid state electrolyte, of the rechargeable battery, e.g., delivered by a mechanical press. The external stress can be applied by a housing, e.g., made of metal. In some cases, the volumetric constraint can be from about 70 MPa to about 1,000 MPa, e.g., about 70 MPa to about 150 MPa, about 100 MPa to about 300 MPa, about 200 MPa to about 400 MPa, about 300 MPa to about 500 MPa, about 400 MPa to about 600 MPa, about 500 MPa to about 700 MPa, about 600 MPa to about 800 MPa, about 700 MPa to about 900 MPa, or about 800 MPa to about 1,000 MPa, e.g., about 70 MPa, about 75 MPa, about 80 MPa, about 85 MPa, about 90 MPa, about 95 MPa, about 100 MPa, about 150 MPa, about 200 MPa, about 250 MPa, about 300 MPa, about 350 MPa, about 400 MPa, about 450 MPa, about 500 MPa, about 550 MPa, about 600 MPa, about 650 MPa, about 700 MPa, about 750 MPa, about 800 MPa about 850 MPa, about 900 MPa, about 950 MPa, or about 1,000 MPa. In the present invention,“about” means ± 10%.
The solid state electrolyte may also be compressed prior to inclusion in the battery. For example, the solid state electrolyte may be compressed with a force between about 70 MPa to about 1,000 MPa, e.g., about 70 MPa to about 150 MPa, about 100 MPa to about 300 MPa, about 200 MPa to about 400 MPa, about 300 MPa to about 500 MPa, about 400 MPa to about 600 MPa, about 500 MPa to about 700 MPa, about 600 MPa to about 800 MPa, about 700 MPa to about 900 MPa, or about 800 MPa to about 1,000 MPa, e.g., about 70 MPa, about 75 MPa, about 80 MPa, about 85 MPa, about 90 MPa, about 95 MPa, about 100 MPa, about 150 MPa, about 200 MPa, about 250 MPa, about 300 MPa, about 350 MPa, about 400 MPa, about 450 MPa, about 500 MPa, about 550 MPa, about 600 MPa, about 650 MPa, about 700 MPa, about 750 MPa, about 800 MPa about 850 MPa, about 900 MPa, about 950 MPa, or about 1,000 MPa. Once pressed, the solid state electrolyte can then be employed in a battery. Such a battery may also be subjected to external stress to enforce a mechanical constriction on the solid state electrolyte, e.g., at the microstructure level, i.e., to provide an isovolumetric constraint. The mechanical constriction on the solid state electrolyte may be from 1 to 100 GPa, e.g., 5 to 50 GPa, such as about 15 GPa. The external stress required to maintain the mechanical constriction may be from about 1 MPa to about 1,000 MPa, e.g., about 1 MPa to about 50 MPa, about 1 MPa to about 250 MPa, about 3 MPa to about 30 MPa, about 30 MPa to about 50 MPa, about 70 MPa to about 150 MPa, about 100 MPa to about 300 MPa, about 200 MPa to about 400 MPa, about 300 MPa to about 500 MPa, about 400 MPa to about 600 MPa, about 500 MPa to about 700 MPa, about 600 MPa to about 800 MPa, about 700 MPa to about 900 MPa, or about 800 MPa to about 1,000 MPa, e.g., about 70 MPa, about 75 MPa, about 80 MPa, about 85 MPa, about 90 MPa, about 95 MPa, about 100 MPa, about 150 MPa, about 200 MPa, about 250 MPa, about 300 MPa, about 350 MPa, about 400 MPa, about 450 MPa, about 500 MPa, about 550 MPa, about 600 MPa, about 650 MPa, about 700 MPa, about 750 MPa, about 800 MPa about 850 MPa, about 900 MPa, about 950 MPa, or about 1,000 MPa. The external stress employed may change depending on the voltage of the battery. For example, a battery operating at 6V may employ an external stress of about 3 MPa to about 30 MPa, and a battery operating at 10V may employ an external stress of about 200 MPa. The invention also provides a method of producing a battery using compression of the solid state electrolyte prior to inclusion in the battery, e.g., with subsequent application of external stress.
Methods
Batteries of the invention may be charged and discharged for a desired number of cycles, e.g., 1 to 10,000 or more. For example, batteries may be cycled 10 to 750 times or at least 50, 100, 200, 300, 400, 500, 600, 700, 800, 900, 1,000, 1,500, 2,000, 3,000, 4,000, or 5,000 times. In embodiments, the voltage of the battery ranges from about 1 to about 20V, e.g., about 1-10V, about 5-10V, or about 5- 8V. Batteries of the invention may also be cycled at any appropriate current density e.g., 1 mA cm-2 to 20 mA cm-2, e.g., about 1-10 mA cm-2, about 3-10 mA cm-2, or about 5-10 mA cm-2.
EXAMPLES
Example 1
The cyclic voltammograms (CV) of Li/LGPS /LGPS+C were measured under different pressures between open circuit voltage (OCV) to 6 V at a scan rate of 0.1mVs-1 on a Solartron electrochemical potentiostat (1470E), using lithium (coated by Li2HPO4) as reference electrode. A liquid battery using LGPS/C thin film as cathode, lithium as anode and, 1 M LiPF6 in EC/DMC as electrolyte was also assembled for comparison. The ratio of LGPS to C is 10:1 in both solid and liquid CV tests.
The cathode and anode thin films used in all-solid-state battery were prepared by mixing
LTO/LCO/LNMO, LGPS, Polytetrafluoroethylene (PTFE) and carbon black with different weight ratios. The ratios of active materials/LGPS/C are 30/60/10, 70/27/3, 70/30/0 for LTO, LCO and LNMO thin film electrodes, respectively. This mixture of powder was then hand-grinded in a mortar for 30 minutes and rolled into a thin film inside an argon-filled glove box with 3% PTFE added. Solid electrolytes used in all-solid-state Li ion batteries were prepared by mixing LGPS and PTFE with a weight ratio of 97:3, then hand-grinding the mixed powder in a mortar for 30 minutes and finally rolling it into a thin film inside an argon-filled glove box. To assemble an all-solid-state Li ion battery cell, the prepared composite cathode (LCO or LNMO) thin film, LGPS thin film (<100 µm), and anode (LTO) thin film were used as cathode, solid electrolyte, and the anode, respectively. The three thin films of cathode, electrolyte and anode were cold-pressed together at 420 MPa, and the pressure was kept at 210 MPa by using a pressurized cell during battery cycling test. The charge and discharge behavior was tested using an ArbinBT2000 workstation (Arbin Instruments, TX, USA) at room temperature. The specific capacity was calculated based on the amount of LTO. Example 2– Strain-Stabilized LGPS Core-Shell Electrolyte Batteries
Theory - The Physical Picture
The mechanism by which strain can expand the LGPS stability window is depicted in Figure 4A. Consider the decomposition of LGPS to some arbitrary set of decomposed products, denoted“?”
Figure imgf000049_0026
, at standard temperature and pressure. The Gibbs energy of the system as a function of the fraction of LGPS that has decomposed is given by the dashed orange line in Figure 4A and
Figure imgf000049_0019
analytically in equation 1.
Figure imgf000049_0001
The lowest Gibbs energy state is
Figure imgf000049_0002
all decomposed) and the initial state is ?? = 0 (pristine LGPS). Accordingly, the reaction energy is
Figure imgf000049_0003
This system is inherently unstable. That is,
Figure imgf000049_0023
is negative for all values of ??. Hence, for any initial value of
Figure imgf000049_0005
the system will move to decrease by increasing ??, ultimately ending at the final state ?
Figure imgf000049_0004
Figure imgf000049_0024
Next, consider the application of a mechanical system that constrains the LGPS particle. Given that LGPS tends to expand during decay, any mechanical constraint will require that decomposition induce strain in the surrounding neighborhood. Such a constraining system could be either materials- level (i.e. a core-shell microstructure) or systems-level (i.e. a pressurized battery cell) or a combination of the two. Ultimately, this mechanical system can only induce a finite strain before fracturing. The energy needed to fracture the system is denoted
Figure imgf000049_0015
Prior to the fracturing of the constraining mechanism, any decomposition of the LGPS must lead to an increase in strain energy. The green line in Figures 5A-5B plots the constrained Gibbs energy (
Figure imgf000049_0016
) terms of the unconstrained Gibbs
Figure imgf000049_0017
and the constraint induced strain The highlighted
Figure imgf000049_0020
curve indicates the decomposition pathway of the LGPS.
1. The particle begins as pristine LGPS
Figure imgf000049_0018
with an unfractured constraint mechanism 2. As the particle begins to decompose
Figure imgf000049_0013
the constraint mechanism requires an increase in The strain Gibbs is assumed to be a function of ?? that goes to zero as
Figure imgf000049_0014
Figure imgf000049_0022
goes to zero
3. Once the Gibbs energy of the strained system exceeds the Gibbs energy of the
Figure imgf000049_0006
fractured system
Figure imgf000049_0012
the constraining mechanism will fail. This occurs at the fracture point
Figure imgf000049_0010
4. Once , the system will proceed to completely decompose as
Figure imgf000049_0011
Figure imgf000049_0021
If the constraint induced strain Gibbs
Figure imgf000049_0025
is sufficiently steep, the slope of the total Gibbs at
Figure imgf000049_0009
?? will be positive (as depicted in Figure 5A). In this case, the LGPS will be metastable about the pristine state
Figure imgf000049_0008
This work focuses on the quantification of constraining systems such that
, allowing metastable ceramic sulfide electrolytes.
Figure imgf000049_0007
Two Work Differentials
The presence of
Figure imgf000050_0012
as a function of ?? stems from the nature of LGPS to expand upon decomposition. Depending on the set of decomposed products, as determined by the applied voltage, this volume expansion can exceed 20 - 50%. As such, the process of LGPS decomposition is one that can include significant“stress-free” strain– that is, strain that is the result of decomposition and not an applied stress. Proper thermodynamic analysis of such decay pathways requires careful consideration of the multiple work differentials, which are reasonably neglected for other systems. Figure 5B schematically represents two sources of work which are frequently used, the“fluid-like” and the“solid-like” forms. In the fluid-like system, the change in work under isobaric conditions is proportional to the change in the system volume
Figure imgf000050_0006
For solid-like systems, the work is defined in terms of a reference/undeformed state and has differential form where
Figure imgf000050_0007
???? is the undeformed volume, ? is the strain tensor relative to the undeformed state and ? is the stress tensor corresponding to ?.
The general approach to showing the equivalency of these two differential work expressions is as follows. The solid-like stress and strain tensors are separated into the compression and distortion terms via the use of deviatoric tensors as defined in equation 2. The pressure is generalized in terms of the stress matrix and volume strain
Figure imgf000050_0013
Figure imgf000050_0008
Figure imgf000050_0001
Using these definitions, the solid-like work can be separated into one term that only includes compression and one term that only includes deformation.
Figure imgf000050_0003
In the fluid limit, where there is no shape change, equation 3 reduces to
Figure imgf000050_0009
assuming that ????? = 0, giving back the fluid-like work differential. In most mechanical systems, this assumption is valid as the undeformed reference volume does not change. However, it fails in describing LGPS decomposition because the undeformed volume changes with respect to ?? and, hence,
Figure imgf000050_0010
Figure imgf000050_0002
Instead, proper thermodynamic analysis of LGPS decomposition requires consideration of both work terms. The fluid term -
Figure imgf000050_0011
indicates the work needed to compress the reference volume (i.e., change ??) in the presence of a stress tensor ? and the solid term represents the work needed to deform the new reference state Considering this, the full energy differential is given by
Figure imgf000050_0005
equation 5.
Figure imgf000050_0004
Transforming to the Gibbs energy
Figure imgf000051_0001
yields the differential form:
Figure imgf000051_0002
Note that the transformation used frequently in solid mechanics,
Figure imgf000051_0003
is sufficient so long as is constant and, hence, can be set as the zero point.
Figure imgf000051_0006
Figure imgf000051_0007
Figure imgf000051_0004
At constant temperature, equation 6 gives the differential form of of Figures 5A-5B in terms of the chemical terms and the strain term
Figure imgf000051_0005
In the following discussion we consider two limiting cases for
Figure imgf000051_0024
?????? as a function of which
Figure imgf000051_0025
provides a range of values for which LGPS can be stabilized. The first case is that of a LGPS particle that decomposes hydrostatically and is a mean field approximation. The fraction of decomposed LGPS is assumed to be uniform throughout the particle
Figure imgf000051_0023
The second limiting case is that of spherically symmetric nucleation, where LGPS is completely decomposed within a spherical region of radius and pristine outside this region
Figure imgf000051_0013
Figure imgf000051_0015
As is shown below, the hydrostatic case yields a lower limit for whereas the nucleation
Figure imgf000051_0014
model shows how this value could, in practice, be much higher.
Hydrostatic Limit/Mean Field Theory
The local stress ? experienced by a subsection of an LGPS particle is directly a function of the decomposition profile
Figure imgf000051_0022
as well as the mechanical properties of the particle and, if applied, the mechanically constraining system. In the hydrostatic approximation, the local stress is said to be compressive and equal everywhere within the particle
Figure imgf000051_0011
In the mean field
approximation, the same is said for the decomposed fraction Given the one-to-one
Figure imgf000051_0012
relation between and these two approximations are equivalent.
Figure imgf000051_0016
Figure imgf000051_0017
We restrict focus to the limit as to evaluate the metastability of LGPS about the pristine state.
Figure imgf000051_0027
If
Figure imgf000051_0018
then the particle is known to be at least metastable with total stability being determined by the magnitude of
Figure imgf000051_0019
The relationship between the pressure and decomposed fraction was shown in r
Figure imgf000051_0021
to be, in this limit, . Where s the effective bulk
Figure imgf000051_0009
Figure imgf000051_0010
modulus of the system, accounting for both the compressibility of the material and the applied mechanical constraint. indicates how much pressure will be required to compress the system
Figure imgf000051_0020
enough as to allow the volume expansion of
Figure imgf000051_0026
that accompanies decomposition. The differential strain Gibbs can be solved from here assuming no deviatoric strain (justifiable for a fluid model) as shown in equation 8.
Figure imgf000051_0008
Figure imgf000052_0012
The reference volume is the volume in the unconstrained system
Figure imgf000052_0001
,
Combining equation 7 and equation 9 with the metastability condition it is found
Figure imgf000052_0009
that fluid-like LGPS will be stabilized whenever equation 10 is satisfied.
Figure imgf000052_0002
Equation 9 is solved for in Figure 5 for the case of a core-shell constriction mechanism with a core comprised of either LGPS or oxygen-doped LGPSO
Figure imgf000052_0010
and a shell of an arbitrary rigid material. The effective bulk modulus is given by
Figure imgf000052_0011
is the compressibility of the LGPS material and is a parameter that represents the ability
Figure imgf000052_0003
of the shell to constrain the particle22.
Spherical Nucleation Limit
The maximally localized (i.e. highest local pressure) decomposition mechanism is that of spherical nucleation as shown in Figure 6. In this model, an LGPS particle of outer radius ?? undergoes a decomposition at its center. The decomposed region corresponds to the material that was initially within a radius of ??. The new reference state is of higher volume than the pristine state as the material has decomposed to a larger volume given by
Figure imgf000052_0015
The decomposed fraction is no-longer a constant in the particle as it was in the hydrostatic case. Instead,
Figure imgf000052_0017
for all material that was initially (prior to decomposition) within the region for all
Figure imgf000052_0016
material initially outside this region,
Figure imgf000052_0005
To fit the decomposed reference state of radius ?? into the void of radius ??, both the decomposed sphere and the remaining LGPS must become strained as shown in Figures 7A.iii and 7A.iv. Thus, solving for the stress in terms of the decomposed fraction ?? becomes the problem of a thick-walled spherical pressure vessel compressing a solid sphere. The pressure-vessel has reference state inner and outer radii given by ?? and ?? and the spherical particle has an equilibrium radius of
Figure imgf000052_0018
Figure imgf000052_0004
In terms of the displacement vector of the decomposed and pristine materials, , and
Figure imgf000052_0006
the radial stress components , the boundary conditions are:
Figure imgf000052_0013
1. Continuity between the decomposed and pristine products:
Figure imgf000052_0007
Where vector notation has been dropped to reflect the radial symmetry of the system.
2. Continuity between the radial components of stress for those materials at the interface
between the decomposed and pristine products:
Figure imgf000052_0008
For a spherically symmetric stress in an isotropic material, the displacement vector is known to be of the form ?
Figure imgf000052_0014
, where the vector notation has been removed as displacement is only a function of distance from the center. The strain Gibbs for a compressed sphere under condition 2, defining
Figure imgf000053_0005
gives the compressive term ?
Figure imgf000053_0006
with no deviatoric components. Likewise, a hollow pressurized sphere at the onset of decay (lim
Figure imgf000053_0007
has both a compressive and deviatoric component that combine to
Figure imgf000053_0002
where ?? is the shear modulus of the pristine material. Combining these terms leads to the nucleated equivalent of equation 8.
Figure imgf000053_0001
Figure 7B shows equation 11 solved for the case where the pristine and decomposed materials have the same elastic modulus (?? = ??) and Poisson’s ratio (?? = ??). The gray and purple lines reflect the no-shell and perfect-shell limits of the hydrostatic model, whereas the blue and red lines represent equation 10 for typical Poisson values. It is seen that, in general, the nucleation model provides a steeper strain Gibbs than the hydrostatic model due to the higher pressures involved. Intuitively, a smaller Poisson’s ratio (harder to compress) improves the stability of the nucleation limit.
Passivation Layer Theory
Electrolytes, either liquid or solid, are likely to react with electrodes where the electrode potential is outside of the electrolyte stability window. To address this, it is suggested that electrolytes be chosen such that they form a passivating solid-electrolyte-interface (SEI) that is at least kinetically stable at the electrode potential. Many works on the topic of improving sulfide electrolytes have speculated that by forming electronically insulating layers on the surface of sulfide electrolytes such passivation layers can be formed. In this section, we discuss the role of such passivation layers and provide a quantitative analysis of the mechanism by which we believe an electronically insulating surface layer improves stability.
In Figure 8A, the thermodynamic equilibrium state is given for the most basic battery half-cell model. A cathode is separated from lithium metal by an electrically insulating and ionically conducting material (? = 0, ? ¹ 0, where ?, ? are the electronic and ionic conductivities) and a voltage ? is applied to the cathode relative to the lithium metal. The voltage of the lithium metal is defined to be the zero point. In terms of the number of electrons (?), the number of lithium ions (?), the Fermi level (ℰ?) and the lithium ion chemical potential (????), the differential Gibbs energy can be written as equation 12 (superscripts ?, ? differentiate the anode from the cathode).
Figure imgf000053_0008
Applying conservation
Figure imgf000053_0003
gives the well-known equilibrium conditions:
Figure imgf000053_0004
Or, in other words, the electrochemical potential (? = ? + ???) of both the electrons and the lithium ions must be constant everywhere within the cell. As a result, the lithium metal potential (??? = ???? + ???) remains constant throughout the cell. The band diagrams found in Figure 7A illustrate how the chemical potential of each species, as well as the voltage, varies throughout the cell, but the electrochemical potential remains constant.
Figure 8B depicts the expected equilibrium state in the case of a solid-electrolyte cathode, where the cathode material is imbedded in a matrix of solid-electrolyte. In this case, the lower (i.e. more- negative) chemical potential of the cathode material relative to the electrolyte causes charge separation that results in an interface voltage ??. Analogous to the procedure following equation 12, it can be shown that the equilibrium points now include the anode (a), cathode (c) and the solid- electrolyte (SE):
Figure imgf000054_0001
Like equation 13, equation 14 leads to the condition that the lithium metal potential remains constant throughout the cell.
A speculated mechanism for passivation layer stabilization of sulfide electrolytes is depicted in Figure 8C. In this case, the solid-electrolyte is coated in an electronically insulating material. Since the external circuitry does not directly contact the solid-electrolyte and there is no electron conducting pathway, the number of electrons within the solid-electrolyte is fixed. Hence the Fermi energy cannot equilibrate via electron flow. The speculation is that this effect could be utilized to allow a deviation of the lithium metal potential within the solid-electrolyte relative to the electrodes, leading to a wider operational voltage window. The band diagrams of Figure 8C illustrate how the electron
electrochemical potential can experience a local maximum (or minimum) in the solid-electrolyte due to a lack of electron conduction. This local maximum (or minimum) is carried over to the lithium metal potential.
The authors believe that while an electronically insulating passivation layer is a key design parameter, the above theory is missing a critical role of effective electron conduction that occurs due to the ‘lithium holes’ that are created when a lithium ion migrates out of the insulated region, leaving behind the corresponding electron. The differential Gibbs energy of this system is represented by adding a solid-electrolyte term to equation 12 (denoted by superscript SE).
Figure imgf000054_0002
The electron and lithium conservation constraints are now:
Figure imgf000054_0003
The effect of removing a lithium ion from the SE is that of placing the
corresponding electron at the Fermi level of the remaining material. 2.
Figure imgf000055_0005
Gaining a lithium ion, but not the corresponding electron, at the anode reduces the number of electrons at the Fermi level.
3. Conservation of total lithium.
Figure imgf000055_0006
Constraints 1 and 2 represent the tethering of the electron and lithium density in the case of an insulated particle. Unlike the system governed by equation 12, the Fermi level of the solid-electrolyte is not fixed by an external voltage. The result is that by lowering the number of atoms within the solid- electrolyte by extracting lithium ions, and hence increasing the number of electrons per atom within the insulated region, the number of electrons per atom and the Fermi level increase. In effect, this represents the conduction of electrons by way of lithium-holes. Solving equation 15 for the equilibrium points given the above constraints lead to those of equation 14 between the anode/cathode as well as the following relation between the anode and solid-electrolyte.
Figure imgf000055_0001
The total voltage experienced within the SE can be represented as where is the
Figure imgf000055_0002
Figure imgf000055_0007
voltage in the absence of lithium extraction from the SE (the original voltage as depicted in Figure 8C) and ?? is the voltage that results from the charge separation of lithium extraction. In other words, the system begins with a charge neutral solid-electrolyte at voltage ???
? . However, equation 16 is not, in general, satisfied. Charge separation occurs lowering the voltage of the solid electrolyte relative to the anode. In terms of a geometrically determined capacitance ?, this charge separation voltage is ?? = ??????? . This effect is illustrated in Figure 8D. Prior to charge separation within the SE region, the voltage and chemical potentials are given by the solid blue lines. As lithium ions are extracted from the SE by the anode, the voltage in the SE decreases from
Figure imgf000055_0003
The ultimate result of this voltage relaxation within the electronically insulated region is depicted in Figure 8E. Because of the effective electron transport via lithium hole conduction, negatively charged lithium metal can form locally within the particle once the applied voltage exceeds the intrinsic stability of the solid-electrolyte. The negative charge is due to the lithium ions that have left the insulated region to equilibrate the lithium metal potential. As such, the local (i.e. within the insulated region) lithium metal is expected to have an interface voltage
Figure imgf000055_0009
with the remaining solid-electrolyte. The voltage must be equal to the voltage between the anode lithium and the solid-electrolyte ?
Figure imgf000055_0004
In short, from a thermodynamic perspective, applying a voltage
Figure imgf000055_0008
to an electronically insulated solid- electrolyte particle relative to a lithium metal anode is equivalent to applying a charged lithium metal directly in contact with the solid-electrolyte.
Intrinsically, this has no impact on the solid-electrolyte stability. However, in the limit of very low capacitances, as is expected, only a small fraction of the lithium ions would need to migrate to the anode for
Figure imgf000055_0010
Hence the electronically insulating shell traps the bulk of the lithium ions locally which maintains the high reaction strain needed for mechanical stabilization. Results and Discussion
Electrochemical Stability
The impact of mechanical constriction on the stability of LGPS was studied by comparing decay metrics between LGPS and the same LGPS with an added core-shell morphology that provides a constriction mechanism. To minimize chemical changes, the constricting core-shell morphology was created using post-synthesis ultrasonication. This core-shell LGPS (“ultra-LGPS” hereafter) was achieved by high-frequency ultrasonication that results in the conversion of the outer layer of LGPS to an amorphous material. Bright-field transmission electron microscopy (TEM) images of the LGPS particles before (Figure 9A) and after (Figure 9C) sonication show the distinct formation of an amorphous layer. Statistically-analyzed energy dispersive X-ray spectroscopy (EDS) (Figures 9B and 9C) shows that this amorphous shell is slightly sulfur deficient whereas the bulk regions of LGPS and ultra-LGPS maintain nearly identical elemental distributions. EDS line-scans on individual [ultra-] LGPS particles (Figures 10-12) confirm that a sulfur-deficient surface layer exists for almost every ultra-LGPS particle whereas no such phenomenon is observed for LGPS particles. Note that this is true for LGPS sonication in both solvents tested, dimethyl carbonate (DMC) and diethyl carbonate (DEC) (Figures 11-13). Simply soaking LGPS in DMC without sonication had no obvious effect (Figure 14). This method of post-synthesis core-shell formation minimizes structural changes to the bulk of the LGPS, allowing us to evaluate the effects of the volume constriction on stability without compositional changes.
The electrochemical stabilities of non-constricted LGPS and constricted ultra-LGPS were evaluated using cyclic-voltammetry (CV) measurements of Li/LGPS/LGPS+C/Ta (Figure 15A) and Li/ultra- LGPS/Ta (Figure 15B) cells respectively, with a lithium reference electrode at a scan rate of 0.1????? and a scan range of 0.5 - 5?. Carbon was introduced here to measure the intrinsic electrochemical stability window of the electrolytes without kinetic compromise.12 For LGPS, oxidation peaks at 2.4? and 3.7? are observed during charging and multiple peaks below 1.6? are observed during discharging. These redox peaks can be attributed to the solid-solid phase transition of Li-S and Ge-S components in LGPS24, confirming that LGPS is unstable and severe decomposition occurred during cycling.
In contrast, the decomposition of ultra-LGPS was largely suppressed, manifested by only one minor oxidation peak at a higher voltage (3V) during charging, and almost no reduction peak during discharging (Figure 15B). In fact, the higher stability of ultra-LGPS is also confirmed by the sensitive electrochemical impedance spectra (EIS) before and after CV tests (Figures 15C, 15D). The EIS shows a typical Nyquist plot of battery-like behavior with charge-transfer semicircles in the medium frequency and a diffusion line in the low frequency. The results show that the total impedance of LGPS composite increased from 300Ω to 620Ω (107% increase) after 3 cycles of CV test (Figure 15C), while that of ultra-LGPS composite only increases by 32% (from 250Ω to 330Ω, Figure 15D). The smaller increase of impedance after cycling indicates that ultra-LGPS is more stable so that less solid phases and grain boundaries are generated due to decomposition. These stability advantages of ultra-LGPS over LGPS were found to be even more prominent when implemented in an all-solid-state half-cell battery. The cycling performance was measured for ???????? (LTO) mixed with carbon and either ultra-LGPS or LGPS as a cathode, ultra-LGPS or LGPS as a separator, and lithium metal as the anode. The cycling performance of each configuration was taken at low (0.02C), medium (0.1C), and high (0.8C) current rates. The results, depicted in Figured 16A- 18B, show that the cycling stability of the ultra-LGPS based half-cells substantially outperforms that of the LGPS based half-cells.
To isolate the decomposition of LGPS in the LTO cathode composite, the solid-electrolyte layers were replaced by a glass fiber separator. Figure 15E shows the charge-discharge profiles of LGPS (LTO+LGPS+C/Glass fiber separator/Li) cycled at 0.5C in the voltage range of 1.0 - 2.2 V. A flat voltage plateau at 1.55^? appeared for 70 cycles, which can be ascribable to the redox of titanium. However, the plateau length decreases from cycle 1 to cycle 70 by almost 85.7%, indicating a large decay of the cathode. On the other hand, ultra-LGPS (LTO+ultra-LGPS+C/Glass fiber separator/Li) (Figure 15F) shows the same flat voltage plateau remaining almost unchanged after 70 cycles. This increase in cathode stability is further confirmed by the cyclic capacity curves (Figures 15G and 15H). For LGPS, the specific charge and discharge capacities decrease from ~159 mAh/g to ~27 mAh/g, and ~170mAh/g to ~28 mAh/g, respectively, after 70 cycle. However, ultra-LGPS demonstrates a much better cyclic stability than its LGPS counterpart. After 70 cycles the discharge capacity is still as high as 160 mAh/g, with only roughly 5% of capacity loss.
In each of these results, those ultra-LGPS particles with core-shell morphologies have outperformed the stability of LGPS counterparts. As discussed in ref22, core-shell designs are proposed to stabilize ceramic-sulfide solid-electrolytes via the volume constraint placed on the core by the shell. This experimental electrochemical stability data agrees with this theory. Sulfur deficient shells, as seen in the case of ultra-LGPS, are expected to lower the effective compressibility of the system and hence increase the volume constraint22. The solid-state half-cell (solid-state cathode + glass fiber/liquid electrolyte + lithium metal anode) performance in the voltage range of 1 - 2.2^? vs lithium
demonstrates that ultra-LGPS has, in practice, improved stability over LGPS in the cases of both LGPS oxidation and reduction. Additionally, the Coulombic efficiency of ultra-LGPS is also higher than that of LGPS, indicating an improved efficiency of charge transfer in the system, and less charge participation in unwanted side reactions.
Decomposition Mechanism
To better understand the mechanism by which LGPS decomposes, TEM analyses were performed to study the microstructure of LTO/[ultra-]LGPS interfaces after cycling. An FIB sample (Figure 19A), in which the composite cathode (LTO+LGPS+C) and separating layer (LGPS) are included, was prepared after 1 charge-discharge cycle versus a lithium metal anode. A platinum layer was deposited onto the cathode layer during FIB sample preparation for protection from ion beam milling. A transit layer with multiple small dark particles exists at the cathode/separator interface (hereafter“LTO/LGPS primary interface), as manifested in the TEM bright-field (BF) images (Figure 19B, Figure 20) and STEM dark-field (DF) images (Figure 19D, Figure 20). The particles within the transit layer of STEM DF images show bright contrast, indicating the accumulation of heavy elements. To understand the chemical composition of this transit layer, STEM EELS (electron energy loss spectroscopy) line-scans were performed. The EELS spectra show that Lik, GeM4,5 (Figures 21A-21B), GeM2,3 and PL2,3 (Figure 15E) peaks exist throughout the transit layer, but sulfur peaks (SL2,3, SL1) only show up inside the bright particles, and are absent in the regions outside the bright particles (EELS spectra 12-14 in Figure 15E). This observation indicates that the bright particles within the transit layer are sulfur-rich, which is not only supported by the bright contrast in STEM image (sulfur is the heaviest element among Li, Ge, P and S), and EELS line-scan observation (Figures 19E, 21A, 21B, 22A, and 22B), but also corroborated by previous studies12 reporting that the decomposition products of LGPS will be sulfur-rich phases including S, LiS, P2S5 and GeS2.
Since the composite cathode layer is composed of LTO, LGPS and C, there will be minor LTO/LGPS interfaces (hereafter“LTO/LGPS secondary interface”) that are ubiquitous within the cathode layer. Figure 19F demonstrates the typical STEM DF image of LTO/LGPS secondary interfaces, in which bright particles with similar morphology show up again. The density of such bright particles is much higher, due to higher carbon concentration within cathode layer and thus facilitated LGPS
decomposition. The corresponding STEM EELS line-scan spectra (Figure 19G) show that strong SL2,3 peaks exist at the interface region, corroborating again that the bright particles are sulfur-rich.
Therefore, sulfur-rich particles exist at both primary and secondary LTO/LGPS interfaces in LGPS half-cells after 1 charge-discharge cycle.
As comparison, Figures 23A-23F show the microstructural and compositional (S)TEM studies for ultra-LGPS half-cells. The primary LTO/ultra-LGPS interface after 1 charge-discharge cycle was characterized by TEM BF image (Figure 23A). A smooth interface was observed between the ultra- LGPS separating layer and the composite cathode layer (Figure 23B). The primary LTO/ultra-LGPS interface is clean and uniform, showing no transit layer or dark particles. The secondary LTO/ultra- LGPS interfaces were also investigated for comparison by STEM DF image, EDS line-scan and EDS mapping (Figures 23C-23E). Results show that the atomic percentage of sulfur continuously decreases, as the STEM EDS line-scan goes from inner ultra-LGPS particle to secondary LTO/ultra- LGPS interface, and finally into LTO+C composite region (Figure 23D and Figures 24A, 24B). In other words, the sulfur-deficient-shell feature of ultra-LGPS particles is maintained after cycling, and no sulfur-rich transit layer is formed at the LTO/ultra-LGPS secondary interface. STEM EDS quantitative analyses (Figure 23F) show that the atomic percentage of sulfur inside ultra-LGPS particle is as high as ~38%, while that of secondary LTO/ultra-LGPS interface is as low as 8%.
These results suggest that the nucleation limit is a more faithful representation of the true decay process than the hydrostatic limit. The sulfur rich particles formed in LGPS have a length scale on the order of ?? » 20^??. In ultra-LGPS, the shell thickness is also roughly
Figure imgf000058_0002
. Hence if we consider the formation of such a sulfur particle near the core-shell boundary in ultra-LGPS, the minimum distance from the center of the sulfur rich particle to the exterior of the shell is
Figure imgf000058_0003
In this case which satisfies the condition needed to apply the nucleated
Figure imgf000058_0004
Figure imgf000058_0001
model. In summary, we know that the LGPS decays via a mechanism that leads to nucleation of sulfur rich particles on the surface. We also know that applying a shell layer with a thickness such that ? » ?? inhibits such decay. These results suggest that the pristine core-shell state is at least metastable with respect to the decay towards the state with nucleated decay just below the core-shell interface.
Conclusions
In summary, we have developed a generalized strain model to show how mechanical constriction, given the nature of LGPS to expand upon decay, can lead to metastability in a significantly expanded voltage range. The precise level to which constriction expands the voltage window is depended on the morphology of the decay. We performed a theoretical analysis of two limits of the decay morphology, the minimally and maximally localized cases. The minimally localized case consisted of a mean field theory where every part of the particle decays simultaneously, whereas the maximally localized case consisted of a nucleated decay. It was demonstrated that, while the maximally localized case was best, both cases had the potential for greatly expanding the stability window. We also developed a theory for the role of an electrically insulating passivation layer in such a stain-stabilized system. This model suggests that such passivation layers aid in stability by keeping lithium ions localized within the particle, maximizing the reaction strain.
Experimental results for the stability performance of LGPS before and after the adding of a constricting shell supports this theory. After the formation of shell via ultrasonication, LGPS demonstrated remarkably improved performance cyclic voltammetry, solid-state battery cycling, and solid-state half-cell cycling. Because the shell was applied in a post-synthesis approach, chemical differences between the core-shell and pure LGPS samples, which might otherwise affect stability, were kept to a minimum. The core-shell is believed to be an instance of mechanically constrained LGPS as during any decomposition, the LGPS core will seek to expand whereas the shell will remain fixed. In order words, the shell provides a quasi-isovolumetric constraint on the core dependent on the biaxial modulus of the shell and the particle geometry.
Analysis of the decay morphology found in LGPS particles but not in ultra-LGPS particle suggests that the nucleated decay limit more accurately reflects the true thermodynamics. It was found that, in LGPS, nucleated sulfur-rich decay centers were embedded in the surface of the LGPS particles after cycling. Further, these nucleated decay centers were not found in the cycled ultra-LGPS. The ultra- LGPS maintained a shell thickness comparable to the decay cites in LGPS (approximately 20^??), which was predicted to be sufficient for the high level of stabilization afforded by the nucleated model. These results, combined with the improved stability of ultra-LGPS, indicate that not only is strain- stabilization occurring, but that the magnitude at which it is occurring is dominated by maximally localized decay mechanism. This is a promising result as such nucleated decay has been shown to provide a larger value of ??????????, opening up the door to solid-state batteries that operate at much higher voltages than what has been reported to date. Methods
Sample preparation
LGPS powder was purchased from MSE Supplies company. Ultra-LGPS was synthesized by soaking LGPS powder into organic electrolytes, such as dimethyl carbonate (DMC) and diethyl carbonate (DEC), and then sonicated for 70h in Q125 Sonicator from Qsonica company, a microprocessor based, programmable ultrasonic processor
Electrochemistry
The cyclic voltammograms (CV) of Li/LGPS/LGPS+C/Ta and Li/ultra-LGPS/ultra-LGPS/Ta cells were measured between 0.5 to 5 V at a scan rate of 0.1mVs-1 on a Solartron electrochemical potentiostat (1470E), using lithium as reference electrode. The electrochemical impedance spectrums of Li/LGPS/LGPS+C/Ta and Li/ultra-LGPS/ultra-LGPS/Ta cells were measured at room temperature both before and after CV tests, by applying a 50 mV amplitude AC potential in a frequency range of 1 MHz to 0.1 Hz. The composite cathode used were prepared by mixing LTO, (ultra-)LGPS, polyvinylidene fluoride (PVDF) and carbon black with a weight ratio of 30:60:5:5. This mixture of powders was then hand-grinded in a mortar for 30 minutes and rolled into a thin film inside an argon- filled glove box. SEs were prepared by mixing (ultra-)LGPS and PVDF with a weight ratio of 95:5, then hand-grinding the mixed powder in a mortar for 30 minutes and finally rolling it into a thin film inside an argon-filled glove box. To assemble a solid-state cell, the prepared composite cathode thin film, (ultra-)LGPS thin film, and Li metal foil were used as cathode, solid electrolyte, and the counter electrode, respectively. The thin films of composite cathode and (ultra-)LGPS were cold-pressed together before assembling into the battery. A piece of glass fiber separator was inserted between (ultra-)LGPS thin film and Li metal foil to avoid interfacial reaction between these two phases. Only 1 drop of 1 M LiPF6 in ethylene carbonate (EC) and dimethyl carbonate (DMC) solution (1:1) was carefully applied onto the glass fiber to allow lithium ion conduction through the separator. Swagelok- type cells were assembled inside an argon-filled glove box. Assembling process of an (ultra-)LGPS battery is the same with that of an (ultra-)LGPS solid-state battery, except that the (ultra-)LGPS SE layer is removed. The charge/discharge behavior was tested using an ArbinBT2000 workstation (Arbin Instruments, TX, USA) at room temperature. The specific capacity was calculated based on the amount of LTO (30 wt%) in the cathode film.
Characterization
For FIB sample preparation, the cold-pressed thin film of composite cathode and (ultra-)LGPS after 1 charge-discharge cycle in (ultra)LGPS solid-state battery was taken out inside an argon-filled glove box. It was then mounted onto a SEM stub and sealed into a plastic bag inside the same glove box. FIB sample preparation was conducted on an FEI Helios 660 dual-beam system. The prepared FIB sample was then immediately transferred into JOEL 2010F for TEM and STEM EDS/EELS characterization. Density functional theory calculations
In order to allow comparability with the Material Project crystal database, all DFT calculations were performed using the Material Project criteria. All calculations were performed in VASP using the recommended Projector Augmented Wave (PAW) pseudopotentials. An energy cutoff of 520 eV with k-point mesh of 1000/atom was used. Compressibility values were found by discretely evaluating the average compressibility of the material between 0 GPa and 1 GPa. Enthalpies were calculated at various pressures by applying external stresses to the stress tensor during relaxation and self- consistent field calculations
Example 3– Computational Method to Select Optimum Interfacial Coating
Like liquid counterparts, the key performance metrics for solid-electrolytes are stability and ionic conductivity. For lithium systems, two very promising families of solid-electrolytes are garnet-type oxides and ceramic sulfides. These families are represented, respectively, by the high-performance electrolytes of LLZO oxide and LSPS sulfide. Oxides tend to maintain good stability in a wide range of voltages but often have lower ionic conductivity
Figure imgf000061_0001
Conversely, the sulfides can reach excellent ionic conductivitie
Figure imgf000061_0002
but tend to decompose when exposed to the conditions needed for battery operation.
Instabilities in solid-electrolytes can arise from either intrinsic material-level bulk decompositions or surface/interfacial reactions when in contact with other materials. At the materials-level, solid- electrolytes tend to be chemically stable (i.e. minimal spontaneous decomposition) but are sensitive to electrochemical reactions with the lithium ion reservoir formed by a battery cell. The voltage stability window defines the range of the lithium chemical potential within which the solid-electrolyte will not electrochemically decompose. The lower limit of the voltage window represents the onset of reduction, or the consumption of lithium ions and the corresponding electrons, whereas the upper limit represents the onset of oxidation, or the production of lithium ions and electrons. The voltage window affects the bulk of any solid-electrolyte particle as the applied voltage is experienced throughout. While interfacial reactions occur between the solid-electrolyte and a second‘coating’ material at the point of contact, these reactions can either be two-bodied chemical reactions, where only the solid- electrolyte and the coating material are reactants, or three-bodied electrochemical reactions, in which the solid-electrolyte, coating material and the lithium ion reservoir all participate. The two types of reactions are state-of-charge or voltage independent and dependent, respectively, as determined by the participation of the lithium ion reservoir.
Prior studies have revealed that the most common lithium ion electrode materials, such as ??????
Figure imgf000061_0003
and
Figure imgf000061_0004
unstable interfaces with most solid electrolytes, particularly the high performance ceramic sulfides. Successful implementation of ceramic sulfides in solid-state batteries may employ suitable coating materials that can mitigate these interfacial instabilities. These coating materials may be both intrinsically electrochemically stable and form electrochemically stable interfaces with the ceramic sulfide in the full voltage range of operation. In addition, if different solid- electrolytes are to be used in different cell components for maximum material-level stability, then the coating materials may also change to maintain chemically stable interfaces.
In short, the choice of a coating material depends on both the type of solid-electrolyte and the intended use of operation voltage (anode film, separator, cathode film, etc.). Pseudo-binary computational methods can approximately solve for the stability of a given interface, but are computationally expensive and have not yet been developed in very-large scale. A major performance bottleneck for high-throughput analysis of interfacial stability has been the cost to construct and evaluate many high-dimensional convex hulls. In the case of material phase stability, the
dimensionality of the problem is governed by the number of elements. For example, calculating the interfacial chemical stability of LSPS and LCO would require a 6-dimensional hull corresponding to the set of elements {Li, Si, P, S, Co, O}. The electrochemical stability of this interface is calculated with the system open to lithium, so that lithium is removed from the set and the required hull becomes 5-dimensional ({Si, P, S, Co, O}).
Here we introduce new computational schemata to more efficiently perform interfacial analysis and hence enable effective high-throughput search for appropriate coating materials given both a solid- electrolyte and an operation voltage range. We demonstrate these schema by applying them to search through over 67,000 material entries from the Materials Project (MP) in order to find suitable coating materials for LSPS, which has shown the highest lithium conductivity of around 25 mS cm-1 , in the cases of both anode and cathode operations. Coating material candidates that are both intrinsically stable at the material level and form stable interfaces with LSPS within the prescribed voltage range are termed“functionally stable.”
To establish standards, we focus on finding anode coating materials which are functionally stable in a window of 0-1.5 volts versus lithium metal and cathode coating materials which are functionally stable in a window of 2-4 volts versus lithium metal. These voltage ranges are based on cycling ranges commonly found in today’s lithium ion batteries. Within the anode range, we are particularly interested in finding materials that are stable at 0 volts versus lithium metal, as it could enable the use of lithium as a commercial anode material.
Due to remaining computational limitations, this work focuses only on those materials that require an LSPS interfacial hull-dimensionality of less than or equal to 8. In other words, materials were only considered if the elements present in that material consisted of {Li, Si, P, S} plus up to four additional elements. A total of 69,640 crystal structures in the MP database were evaluated for material-level voltage windows. Of those, 67,062 materials satisfied the less than 8-dimesional requirement and were accordingly evaluated for functional stability with LSPS. In total, over 1,000 MP entries were found to be functionally stable in the anode range and over 2,000 were functionally stable in the cathode range for LSPS. Experimental probing of interfacial stability is used for select materials to confirm these predictions. Results and Discussion
Data Acquisition and Computational Efficiency
To efficiently evaluate the stability of the interface between each of these 67,062 potential coating materials and LSPS, two new computational schemata were developed. To minimize the number of hulls that must be calculated, the coating materials were binned based on elemental composition. Each unique set of elements requires a different hull, but elemental subsets can be simultaneously solved. For example, the calculation of interfacial stability between LSPS and iron-sulfate (??? ???? ? ?) requires solving for the convex hull of the 6-dimensional element set {Li, Si, P, S, Fe, O}. This hull is the same hull that must be calculated for the interface with LFPO and includes, as a subset, the 5- dimensional hull needed for the evaluation of iron-sulfide (???). To capitalize on this, rather than iterate through each of the 67,062 materials and calculate the hull needed for that material, the minimum number of elemental sets that spans the entirety of the materials were determined (Figure 25A). Then for each elemental set, only one hull is needed to evaluate all of materials that can be constructed using those elements. This approach reduces the total number of hulls needed from 67,062 (one per material) to 11,935 (one per elemental set). As seen in Figure 25A, few hulls with a dimensionality below 7 were needed. Those compounds that would otherwise require a low dimensional hull are solved as a subset of a larger element set. Additionally, the number of required 7 and 8 dimensional hulls are largely reduced due to multiple phases of the same compositional space requiring the same hull.
The second schema used to minimize computational cost was a binary search algorithm for determining the pseudo-binary once a hull was calculated. The pseudo-binary approach is illustrated in Figure 25B. Since decomposition at an interface between two materials can consume an arbitrary amount of each material, the fraction of one of the two materials (? in equation 1) consumed can vary from 0-1.
Figure imgf000063_0001
The pseudo-binary is a computational approach that determines for which value of ? the
decomposition described by equation 1 is the most kinetically driven (e.g. when is the decomposition energy the most severe). The RHS of equation 1 represents the fraction ({??}) of each of the thermodynamically favored decay products and defines the convex hull for a given ? in terms of the products’ Gibbs energies The total decomposition energy accompanying
Figure imgf000063_0004
equation 1 is:
Figure imgf000063_0002
The most kinetically driven reaction between LSPS and the coating material is the one that maximizes the magnitude (i.e. most negative) of equation 2, which defines the parameter ??.
Figure imgf000063_0003
This maximum decomposition energy is the result of two factors. The first, denoted , is the portion
Figure imgf000064_0011
of the decomposition energy that is due to the intrinsic instability of the two materials. In terms of the decomposed products of LSPS
Figure imgf000064_0013
and the coating material
Figure imgf000064_0004
? is the decomposition energy corresponding to the reaction
Figure imgf000064_0003
By subtracting this materials-level instability from the total hull energy, the effects of the interface can be isolated
Figure imgf000064_0012
as defined in equation 4.
Figure imgf000064_0001
Physically, represents the instability of the materials when separated and represents
Figure imgf000064_0002
Figure imgf000064_0010
the increase in instability caused by the interface once the materials are brought into contact.
In this work, to determine the added instability of each interface at the most kinetically driven fraction
Figure imgf000064_0015
we implement a binary search algorithm (see Methods) that uses the concavity of the hull to find to within 0.01% error. This binary search approach finds the ?? value in 14 steps of hull evaluations. A more traditional linear evaluation of the hull to 0.01% accuracy would require 10,000 equally spaced evaluations from ? This increase of speed is leveraged to efficiently
Figure imgf000064_0009
search the 67,062 material entries for functional stability.
Functional Stability
Functional stability at a given voltage was determined for each of the 67,062 materials by requiring that (i) the material’s intrinsic electrochemical stability per atom at that voltage was below thermal energy
Figure imgf000064_0005
and (ii) that the added interfacial instability at the given voltage was below thermal energy Under these conditions, the only instability in the system is
Figure imgf000064_0006
that of the LSPS intrinsic material-level instability, which can be stabilized via strain induced methods22. Of the 67k materials, 1,053 were found to be functionally stable in the anode range (0-1.5 V vs. lithium metal) and 2,669 were found to be functionally stable in cathode range (2-4 V vs. lithium metal). Additionally, 152 materials in the anode range and 142 materials in the cathode range were determined to violate condition (i) but only decompose by lithiation/delithation. The practical use of such materials as an LSPS coating material depends on the reversibility of this lithiation/delithiation process, as such these materials are referred to as potentially functionally stable. All functionally stable and potentially functionally stable materials are cataloged in the supplementary information and indexed by the corresponding Materials Project (MP) id.
The correlation between each element’s atomic fraction and the interfacial stability is depicted in Figure 25C and Figures 26A-26C. Figure 25C depicts the correlation of each element with
Figure imgf000064_0014
for chemical reactions whereas Figures 26A-26C depict the correlations with
Figure imgf000064_0008
for electrochemical reactions at 0, 2 and 4 V versus lithium metal, respectively. A negative correlation between elemental composition and implies that increasing the content of that element
Figure imgf000064_0007
improves the interfacial stability. Figure 25C indicates that chemical stability is best for those compounds that contain large anions such as sulfur, selenium and iodine. In general, Figures 26A and 26C indicate that there is reduced correlation between elemental species and at low and high voltages, respectively. This suggests that at these voltage extremes, the interfacial decomposition is dominated by intrinsic materials-level reduction/oxidation
Figure imgf000065_0003
rather than interfacial effects
Figure imgf000065_0002
lithium (Figure 26B) positive correlation (higher instability) is seen for most elements with the notable exception of the chalcogen and halogen anion groups, which are negatively correlated.
Anionic Species Impact on Material-Level Stability
Given the high correlation contrast for anionic species with respect to interfacial stability, analysis of the dataset in terms of anionic composition was performed. To eliminate overlap between the datapoints, the only compounds that were considered were those that are either monoanionic with only one of {N, P, O, S, Se, F, I} or oxy-anionic with oxygen plus one of {N, S, P}.45,580 MP entries met one of these criteria as is outlined in Table 3. The percentage of each anionic class that was found to be electrochemically stable at the material-level is also provided.
Table 3. Sizes of monoanionic and oxy-anionic datasets and the percentage of each that is electrochemically stable in the anode range (0-1.5V) and the cathode range (2-4V). For example, F represents all compounds that contain F in the chemical formula, while O+N represents all compounds that contain both O and N in the chemical formula.
Figure imgf000065_0001
Figure 27A illustrates the impact of applied voltage on the hull energy of a material, in this case LSPS. When the slope of the hull energy with respect to voltage is negative, the corresponding
decomposition is a reduction, whereas it is an oxidation if the slope is positive. In the middle there is a region where the hull slope is zero, implying there is no reaction with the lithium ion reservoir (i.e. the reaction is neutral with respect to lithium). Considering this, Figures 27B and 27C plot the
characteristic redox behavior of each anionic class in the anode and cathode ranges, respectively. The“neutral decay” line at 45 represents those compounds that have the same hull energy at both voltage extremes and hence aren’t reacting with the lithium ions. Datapoints above [below] this line are increasing [decreasing] in hull energy with respect to voltage and are hence are characteristically oxidative [reductive] in the plotted voltage range. Figure 27B indicates that, in agreement with expectations, most compounds are reduced in the anode voltage range of 0-1.5 V vs. lithium metal. Nitrogen containing compounds are seen to
disproportionately occupy the y-axis, indicating a higher level of stability when in direct contact with lithium metal. This is in line with prior computation work that indicates binary and ternary nitrides are more stable against lithium metal than sulfides or oxides33. Within the cathode voltage range (Figure 27C), however, much more variance in anionic classes is seen. The oxy-anionic and fluorine containing compounds remain principally reductive whereas the phosphorous, sulfide, and selenium containing compounds are characteristically oxidative. Oxygen containing compounds are found on both side of the neutral decay line, implying that oxides are likely to lithiate/delithiate in this 2-4V range.
The average hull energy of each anionic class is given in 0.5V steps from 0-5V in Figure 27D.
Nitrogen containing compounds are confirmed to be the most stable at 0V with iodine and phosphorous compounds maintaining comparable stability. Phosphorous and iodine surpass nitrogen in average stability for voltages above 0.5V and 1.0V, respectively. At high voltages (>4V), it is seen that fluorine and iodine containing compounds are stable whereas nitrogen containing compounds are the least stable.
Anionic Species Impact on Interface-Level Stability
The average values of total decomposition energy
Figure imgf000066_0002
and the fraction that is a result of the interface instability
Figure imgf000066_0001
are depicted in Figures 28A-28C for each anionic class. Figure 28A shows the average instability due to chemical reactions between the anionic classes and LSPS. Sulfur and selenium containing compounds form, on average, the most chemically inert interfaces with LSPS. Conversely, fluorine and oxygen containing compounds are the most reactive. As a general trend, those compound classes that are more unstable in total terms (higher ????? ??? ?) also maintain a higher interfacial contribution
Figure imgf000066_0004
relative to the intrinsic material contribution
This implies that the difference of each class’s intrinsic chemical stability plays a less
Figure imgf000066_0003
significant role than its reactivity with LSPS in determining the chemical stability of the interface. Figure 28B shows the average total electrochemical decomposition energy for the interfaces in 0.5V steps from 0-5V. In general, each anionic class follows a path that appears to be dominated by the materials-level electrochemical stability of LSPS (Figure 27A). This is particularly true in the low voltage (<1V) and high voltage (>4V) regimes, where electrochemical effects will be the most pronounced. The biggest deviations of the interfacial stability from LSPS’s intrinsic stability occur in the region of 1-3V. Those compounds with the lowest chemical decomposition energies (compounds containing S, Se, I, P) deviate the least from LSPS within this‘middle’ voltage range, while those with large decomposition energies (compounds containing N, F, O, O+) deviate more significantly. This trend suggests that the low and high voltage ranges are dominated by materials-level electrochemical reduction and oxidation, respectively, while the middle range is dominated by interface-level chemical reactions. For example, at 0V the interface between ????? and LSPS is expected to decay to {??????, ????, ????, ????, ???????} which is the same set of decay products that would result from each material independently decomposing at 0V. Hence the existence of the interface has no energetic effect.
The average interface-level contribution for electrochemical decomposition is shown in Figure 28C. All anionic classes trend to ??
???? ???? = 0 at 0V, implying that the materials tend to become fully reduced at 0V, in which case interfacial effects are negligible compared to material-level instabilities.
Significant interfacial instabilities arise in the middle voltage range and lower again in the high voltages. Again, this implies that interface-level chemical effects are dominant in the middle voltage range whereas material-level reduction [oxidation] dominate at low [high] voltages. At high voltage, the interfacial contribution to the instability approaches the reaction energy between the maximally oxidized material and LSPS. As a result, for any voltage above 4V, the interface will add an instability of energy equal to this chemical reaction. This explains the high-voltage asymptotic behavior, whereas the low-voltage behavior always trends towards 0 eV atom-1. For example, for any voltage above 4V, LFPO will decompose to {??, ?????} whereas LSPS will decompose to
Figure imgf000067_0001
The introduction of the interface allows these oxidized products to chemically react and form
Figure imgf000067_0002
and ????.
Anionic Species Impact of Functional Stability
The total number of each anionic class that were determined to be functionally stable or potentially functionally stable are given in Figure 29A (anode range) and Figure 29B (cathode range), where they are both intrinsically stable at the material level and form stable interfaces with LSPS within the prescribed voltage range. For the anode range, nitrogen, phosphorous, and iodine containing compounds have the highest percentage of stable compounds (2-4%), whereas all other classes are below 1%. The cathode range showed much higher percentages with sulfur containing compounds reaching 35%. Iodine and selenium were both above 10%.
Experimental Comparison
The chemical compatibility between various coating materials and LSPS were tested experimentally by hand-milling the mixture powder of LSPS and coating materials with/without high-temperature annealing, followed by X-ray diffraction (XRD) measurements at room temperature. Any chemical reaction between the powder will cause compositional and structural changes in the original phases, which can be detected by the change of peak positions and intensities in XRD patterns. It is worth noting that even interfacial reactions are predicted to happen based on thermodynamic calculations, a certain amount of energy may be needed to overcome the kinetic energy barrier for these reactions to happen4. Therefore, the mixed powders were annealed at high temperatures (300°C, 400°C, 500°C) to determine the onset temperature of interfacial reactions as well as the reaction products, and to further assess the role of kinetics by comparing these results with the DFT computed thermodynamic reaction products.
Figures 30A-30D compares the XRD patterns of such room-temperature and 500°C-annealed powder mixtures. Several candidate coating materials (i.e. SnO2, Li4Ti5O12, SiO2) were mixed with LSPS (Figures 30C-30D), while the mixed powder of LCO+LSPS was for comparison (Figure 30A). The XRD patterns for each individual phase (i.e. SnO2, Li4Ti5O12, LiCoO2, SiO2 and LSPS) at room temperature and 500°C are used as reference (Figures 31A-31E). By comparing these XRD patterns, it is obvious that at room temperature, no coating materials reacts with LSPS, since the XRD patterns only show peaks of the original phases. However, after being annealed at 500°C for 6h, different materials show completely different reaction capabilities with LSPS. LCO is observed to react severely with LSPS, because the peak intensities and positions of the XRD pattern for the mixed powders changed completely in the whole 2-theta range of 10-80 degrees (Figure309A). The original LCO and LSPS peaks either disappeared or decreased, while extra peaks belonging to new reaction products appeared (such as SiO2, Li3PO4, cubic Co4S3 and monoclinic Co4S3), indicating that LCO is not compatible with LSPS. As a sharp contrast, peak intensities and positions of the XRD patterns for SiO2+LSPS mixture never change, showing only original peaks both before and after 500°C annealing. This is the direct evidence to show that no interfacial reaction happens when SiO2 is in contact with LSPS, despite large external energy provided. SnO2 and LTO also show incompatibility with LSPS, as new peaks belonging to reaction products appeared in the XRD patterns for their 500°C-annealed sample, however, the peaks of reaction products are much weaker than the case of LCO+LSPS. The 2-theta ranges, where peak positions and intensities change for four materials, are highlighted by color regions in Figures 30A-30D, as an indication of the incompatibility of different materials with LSPS. It can be observed from Figures 30A-30D that such incompatibility order is LCO>SnO2>LTO>SiO2, which is in perfect agreement with our theoretical prediction based on thermodynamic calculations. The onset temperature for interfacial reactions of various materials with LSPS are shown in Figures 32A-32D.
The electrochemical stability of typical coating materials is characterized by Cyclic Voltammetry (CV) technique, in which the decomposition of the tested coating material can be manifested by current peaks at certain voltages relevant to Lithium. Two typical coating materials were used as a demonstration to show good correspondence between our theoretical prediction and experimental observation. The CV test of Li2S (Figure 30E) shows a relevantly flat region between 0-1.5V, while a large oxidation peak dominates the region of 2-4V. In contrast, the CV test of SiO2 (Figure 30F) demonstrates net reduction in the region of 0-1.5V, and a neutral region with little decomposition between 2 and 4V. These results are again direct evidence to corroborate our theoretical predictions based on thermodynamic calculations.
Methods
Data Acquisition
The data used in this work was the result of prior Density Functional Theory calculations that were performed as part of the Materials Project (MP) and was interfaced with using the Materials
Application Programming Interface (API). The Python Materials Genomics (pymatgen) library was used to calculate convex hulls. Of the initial 69,640 structures that were evaluated, 2,578 structures were not considered due to requiring hulls of dimension equal to or greater than 9. Elemental Set Iterations
To minimize the computational cost of analyzing all 67,062 structures, the smallest number of elemental sets that spanned all the materials were determined. To do this, the set of elements in each structure were combined with the elements of LSPS, resulting in a list of element sets with each set’s length equal to the dimensionality of the required hull for that material. This list was ordered based on decreasing length of the set (e.g. ordered in decreasing dimensionality of the required hull). This set was then iterated through and any set that equals to or is a subset of a previous set was removed. The result was the minimum number of elemental sets, in which every material could be described. Chemical decomposition hulls were calculated using the energies and compositions from the MP. Changes in the volume and entropy were neglected (D? » D?). Similarly, electrochemical decomposition hulls were founded by using the lithium grand canonical free energy and subtracting a term ?????? from the energies (DF » D? - ???D???), where ??? is the chemical potential of interest and ??? is the number of lithium ions in the structure. After a hull was calculated, it was used to evaluate every material that exists within the span of its elemental set.
The Pseudo-binary
The pseudo-binary, as described in section 2, seeks to find the ratio of LSPS to coating material such that the decomposition energy is the most severe and, hence, is the most kinetically driven. This problem is simplified by using a vector notation to represent a given composition by mapping atomic occupation to a vector element. For example,
Figure imgf000069_0005
in the basis of (Li Co O), meaning that there are 1 lithium, 1 cobalt, and 2 oxygen in the unit formula. Using this notation, the decomposition in equation 1 can be written in vector form.
Figure imgf000069_0001
Using ? to represent a vector and
Figure imgf000069_0006
to represent a matrix, equation 5 becomes:
Figure imgf000069_0002
The relative composition derivatives for each decay product can be found by inverting
Figure imgf000069_0007
in equation 6.
Figure imgf000069_0003
Equation 7 allows for the calculation of the derivative of the hull energy with respect to the fraction parameter ?.
Figure imgf000069_0004
By using equation 7, and the fact that the hull is a convex function of a binary search can be performed to find the maximum value of ????? and the value at which it occurs
Figure imgf000070_0007
? This process consists of first defining a two-element vector that defines the range in which is known to exist
Figure imgf000070_0006
?????? = ?0,1? and an initial guess
Figure imgf000070_0004
? Evaluating the convex hull at the initial guess yields the decomposition products
Figure imgf000070_0003
} and the corresponding energies
Figure imgf000070_0008
. Equations 7 and 8 can then be used to find the slope of the hull energy. If the hull energy is positive,
Figure imgf000070_0002
, whereas if it is negative
Figure imgf000070_0005
This process is repeated until the upper and lower limits differ by a factor less than the prescribed threshold of 0.01%, which will always be achieved in 14 steps (
Figure imgf000070_0009
0.006%?. Equations 5-8 are defined for chemical stability. In the case of electrochemical (lithium open) stability, the free energy is replaced with F
Figure imgf000070_0001
where ? is the chemical potential and ?? is the number of lithium in structure ?. Additionally, lithium composition is not included in the composition vectors of equation 6 to allow for the number of lithium atoms to change.
X-ray Diffraction
The compatibility of the candidate materials and solid electrolyte was investigated at room
temperature (RT) by XRD. The XRD sample was prepared by hand-milling the candidate materials (LCO, SnO2, SiO2, LTO) with LSPS powder (weight ratio=55:30) in an Ar-filled glovebox. To test the onset temperature of reactions for candidate materials and LSPS solid electrolyte, the powder mixtures were well spread on a hotplate to heat to different nominal temperatures (300,400 and 500 degree Celsius) and then characterized by XRD.
XRD tests were performed on Rigaku Miniflex 600 diffractometer, equipped with Cu Ka radiation in the 2-theta range of 10-80°. All XRD sample holders were sealed with Kapton film in Ar-filled glovebox to avoid air exposure during the test.
Cyclic Voltammetry
Candidate coating materials (Li2S and SiO2), carbon black, and poly(tetra- fluoroethylene) (PTFE) were mixed together in a weight ratio of 90:5:5 and hand-milled in an Ar-filled glovebox. The powder mixtures were sequentially hand-rolled into a thin film, out of which circular disks (5/16-inch in diameter, ~1-2 mg loading) were punched out to form the working electrode for Cyclic
Voltammetry(CV) test. These electrodes were assembled into Swagelok cells with Li metal as the counter electrode, two glass fiber separators and commercial electrolyte (1M LiPF6 in 1:1 (volumetric ratio) ethylene carbonate/dimethyl carbonate (EC/DMC) solvent).
CV tests were conducted by Solartron 1455A with a voltage sweeping rate of 0.1mV/s in the range of 0-5V at room temperature, to investigate the electrochemical stability window of the candidate coating materials (Li2S and SiO2). Conclusion
Our high-throughput pseudo-binary analysis of Material Project DFT data has revealed that interfaces with LSPS decay via dominantly chemical means within the range of 1.5 to 3.5 V and electrochemical reduction [oxidation] at lower [higher] voltages. The fraction of decomposition energy attributed to interfacial effects disappears as the voltage approaches 0V. This result suggests that all material classes tend to decay to maximally lithiated Li binary and elemental compounds at low voltage, in which case the presence of the interface has no impact.
In terms of anionic content, we see that appropriately matching operational conditions to the coating material is paramount. Sulfur and selenium containing compounds, for example, demonstrate a very high chance to be functionally stable (>25% among all sulfides and selenides) in the 2-4V cathode range. However, less than 1% of these same materials form a functionally stable coating material in the 0-1.5V anode range, where iodine, phosphorous and nitrogen have the highest performance. Oxygen containing compounds have a high number of phases that are functionally stable in both voltage regions, but the percentage is low due to the even higher number of oxygen containing datapoints.
Example 4
We show that an advanced mechanical constriction method can improve the stability of lithium metal anode in solid state batteries with LGPS as the electrolyte. More importantly, we demonstrate that there is no Li dendrite formation and penetration even after a high rate test at 10 mA cm-2 in a symmetric battery. The mechanical constriction method is technically realized through applying an external pressure of 100 MPa to 250 MPa on the battery cell, where the Li metal anode is covered by a graphite film (G) that separates the LGPS electrolyte layer in the battery assembly. At the optimal Li/G capacity ratio, it exhibits excellent cyclic performances in both Li/G-LGPS-G/Li symmetric batteries and Li/G-LGPS-LiCoO2 (LiNbO3 coated) batteries. Upon cycling, Li/G anode transforms from two layers into one integrated composite layer. Comparison between Density Functional Theory (DFT) data and X-ray Photoelectron Spectroscopy (XPS) analysis yields the first ever direct observation of mechanical constriction controlling the decomposition reaction of LGPS. Moreover, the degree of decomposition is seen to become significantly suppressed under optimum constriction conditions.
Design of Li/Graphite anode
We first investigated the chemical stability between LGPS and (lithiated) graphite through the high temperature treatment of their mixtures at 500℃ for 36 hours inside the argon filled glovebox for an accelerated reaction. XRD measurements were performed on different mixtures before and after heat treatment, as shown in Figures 33(A, B, C). Severe decomposition of LGPS in contact with lithium was observed accompanied with Li2S, GeS2 and Li5GeP3 formation (Fig.33A). In contrast, no peak change occurred for the mixture of LGPS and graphite after heating, as shown in Fig.33B, demonstrating that graphite was chemically stable with LGPS. After heating the mixture of Li and graphite powders, lithiated graphite was synthesized (Fig.38). When the lithiated graphite was further mixed with LGPS, it was chemically stable as shown in Fig.33C, with only a slight intensity change for the 26o peak.
The Li/graphite anode was designed as shown in Fig.33(D). The protective graphite film was made by mixing graphite powder with PTFE and then covering onto the lithium metal. The three layers of Li/graphite, electrolyte and cathode film were stacked together sequentially, followed by a mechanical press. The pressure was maintained at 100-250 MPa during the battery test. Such pressure helps obtain a good contact between anode and electrolyte based on the conventional wisdom in this field, but more importantly, it serves a mechanical constriction for improved electrochemical stability of solid electrolyte. Scanning electron microscopy (SEM) shows that the graphite particles transform into a dense layer under such high pressure (Fig.39). The as-prepared anode before battery test can be directly observed via SEM and focused ion beam (FIB)-SEM in Fig.33E, 33F). The three layers of Li, graphite and LGPS were clear with close interface contact.
Cyclic and rate performance of Li/Graphite anode
The electrochemical stability and rate capability of Li/graphite (Li/G) anode was tested with anode- LGPS-anode symmetric battery design under 100 MPa external pressure. The comparison of cyclic performance between Li/G-LGPS-G/Li and Li-LGPS-Li batteries is shown in Fig.34A. Li symmetric battery works only for 10 hours at a current density of 0.25 mA cm-2 before failure, while Li/G symmetric battery was still running after 500 hours of cycling with the overpotential increasing slowly to 0.28 V. The stable cyclic performance was repeatable, as shown in Fig.40 from another battery with a slower overpotential increase from 0.13 V to 0.19 V after 300 hours’ cycling, indicating such slight overpotential change varies from battery assembly. SEM shows that Li/Graphite anode transforms from two layers to one integrated layer of composite without notable change of total thickness after long-term cycling (Fig.41). The SEM images of Li/G anode after 300 hours’ cycling in a symmetric battery were compared with the Li anode after 10 hours’ cycling in Fig.34B. The Li/G anode maintained a dense layer of lithium/graphite composite after the long-term cycling (Fig.34B1, B2). In comparison, countless pores appeared in the Li anode after 10 hours of test, which were most probably induced by severe decomposition reaction of LGPS with Li metal. The pores were harmful to both ionic and electronic conductivities, which might be responsible for the sharp voltage increase when Li symmetric battery fails at 10 hours.
We also compared the rate performance of Li/G symmetric battery under different external pressures of 100 MPa or 3 MPa as shown in Fig.34C. Same charging and discharging capacities were set for different current densities by changing the working time per cycle. The Li/G symmetric battery can cycle stably from 0.25 mA cm-2 up to 3 mA cm-2 with an overpotential increase from 0.1 V to 0.4 V. It can then cycle back normally to 0.25 mA cm-2 (Fig.34C1). While at 3 MPa, the battery failed during the test at 2 mA cm-2 (Fig.34C2). Note that at the same current density, the overpotential at 100 MPa was only around 63 % of that under 3 MPa. The SEM images of the Li/G-LGPS interface after the rate test up to 2 mA cm-2 showed a close interface contact at 100 MPa (Fig.34D1), while cracks and voids were observed after the test at 3 MPa (Fig.34D2). Thus, the external pressure plays the role of maintaining the close interface contact during the battery test, contributing to the better rate performance.
To further understand the influence of the Li/G composite formed by battery cycling on its high rate performance, a battery test was designed like Fig.34(E1). Here, a higher external pressure of 250 MPa was kept during the test. It starts at 0.25 mA cm-2 for 1 cycle and then directly goes to 5 mA cm-2 charge, which shows a sharply increased voltage that leads to the safety stop. We then restarted the battery instantly, running at 0.25 mA cm-2 again for ten cycles followed by 5 mA cm-2 for the next ten. This time the battery runs normally at 5 mA cm-2 with an average overpotential of 0.6 V, and it can still go back to cycle at 0.25 mA cm-2 without obvious overpotential increase. At fixed current, the initial voltage surge at 5 mA cm-2 indicates a resistance jump, which is most probably related to the fact that Li and graphite are two layers as assembled, and hence there is not sufficient Li in graphite to support such a high current density. However, after 20 hours’ cycling at 0.25 mA cm-2, Li/G was on the track of turning into a composite, as shown in Fig.34B and Fig 41, with much more Li storage to support the high rate cycling test.
Based on the above understanding, we further lowered the current density for the initial cycles to 0.125 mA cm-2 and cycled with the same capacity of 0.25 mAh cm-2 for a more homogeneous Li distribution and storage in the Li/G composite for improved lithium transfer kinetics. As shown in Fig. 34(E2), the battery could cycle at a current density of 10 mA cm-2 and cycle normally when the current density was set back to 0.25 mA cm-2. Note that there was no obvious overpotential increase at the same low current rate before and after the high rate test, as shown in the insets of Fig.34E and Fig. 42, where the SEM of Li/G anode of this battery also showed a clear formation of Li/G composite without obvious Li dendrite observed on the interface.
Li/Graphite anode in all-solid-state battery
We first performed DFT simulations of LGPS decomposition pathways in the low voltage range of 0.0- 2.2V versus lithium metal. Mechanical constriction on the materials level was parameterized by an effective bulk modulus (Keff) of the system. Based on the value of this modulus, the system could range from isobaric (Keff = 0) to isovolumetric (Keff = ¥). Expected values of Keff in real battery systems were on the order of 15GPa. In the following, these simulation results were used to interpret XPS results of the valence changes of Ge and P from LGPS in the solid state batteries after CV, rate and cycling tests.
As shown in Figure 36A, the decomposition capacity of LGPS was lower at high effective moduli, indicating that the decomposition of LGPS at low voltage was largely inhibited by mechanical constriction. The predicted decomposition products and fraction number are listed in Figure 36B and Table 4, respectively. At Keff = 0 GPa (i.e. no applied mechanical constraint/isobaric), the reduction products approached the lithium binaries Li2S, Li3P, and Li15Ge4 as the voltage approaches zero. However, after mechanical constriction was applied and the effective modulus was set at 15 GPa, the formation of Ge element, LixPy and LixGey were suppressed, while compounds like PxGey, GeS, and P2S were emergent. This is also in agreement with the fact that PxGey is known to be a high pressure phase. The voltage profiles and reduction products at different Keff shown in Figure 36 indicate that the decomposition of LGPS follows different reduction pathways at low voltage after the application of mechanical constriction.
Table 4. (A)-(D) LGPS decomposition products with fraction numbers down to low voltages at different Keff
(A) Keff = 0 GPa
Figure imgf000074_0001
Figure imgf000075_0001
Figure imgf000076_0001
Figure imgf000077_0001
It is worth noting that while the applied pressure and the effective modulus (Keff) were both measured in units of pressure, they are independent. The effective modulus represents the intrinsic bulk modulus of the electrolyte added in parallel with the finite rigidity of the battery system. Accordingly, Keff measures the mechanical constriction that can be realized on the materials level in any single particle, while the external pressure applied on the operation of solid state battery enforced the effectiveness of such constriction on the interface between particles or between electrode and electrolyte layers. This is because exposed surface was the most vulnerable to chemical and electrochemical decompositions, while a close interface contact enforced by external pressure will minimize such surface. Thus, even though the applied pressure was only on the order of 100 MPa, the effective bulk modulus was expected to be much larger. In-fact, close packed LGPS particles should experience a Keff of approximately 15GPa. The applied pressure of 100-250 MPa was an effective tool for obtaining this close packed structure. In short, the applied pressure minimizes gaps in the bulk electrolyte, allowing for the effective modulus that represents the mechanical constriction on the materials level to approach its ideal value of circa 15 GPa.
The XPS results of LGPS that was either in direct contact with a lithium or lithium-graphite anode, as well as bulk LGPS during battery cycling are provided in Figure 37. These measurements of valence change can be well understood in light of the phase predictions of Figure 36B. LGPS in the separator region far from the anode interface showed Ge and P peaks identical to the pristine LGPS (Fig.37A). We first investigate the function of Li/G composite in comparison with pure lithium metal at a slow rate of 0.25 mA/cm2 under 100 MPa external pressure (Fig.37B, C). With pure lithium metal (Fig.37C) the reductions of both Ge and P were significant on the Li-LGPS interface, showing the formation of LixGey alloy, elemental Ge, and Li3P. Note that Ge valence in LixGey and P valence in Li3P are negative or below zero valence, consistent with the Bader charge analysis from DFT simulations (Fig. 44.) In contrast, with the Li/G anode the reductions were inhibited on the Li/G-LGPS interface, with both Ge and P valences remaining above zero in the decomposed compounds (Fig.37B). The Li and LGPS interface was chemically unstable, leading to decompositions that include the observed compounds in Fig.37C. These decompositions were also consistent with the predicted ones in Fig. 36B at Keff at 0 GPa. Further electrochemical cycling of such chemically decomposed interface will cause the decomposed volume fraction to grow, ultimately consuming all of the LGPS. On the contrary, graphite layer in Li/G anode prevented the chemical interface reaction between LGPS and Li, while under proper mechanical constriction the electrochemical decomposition seems to go through a pathway of high Keff ³ 10 GPa in Fig.36B, where GeS, PxGe, P2S match the observed valences from XPS in Fig.37B.
When the cycle rate was increased to 2 mA/cm2 and 10 mA/cm2, the observed decompositions on the L/G-LGPS interface under external pressures in Fig.37D, 37E changed to a metastable pathway that was different from the low rate one at 0.25 mA/cm2 in Fig.37B. This implies that while Fig.37B agrees with the thermodynamics predicted in Figure 36, at high current densities the decomposition becomes kinetically dominated. Moreover, it was concluded that the Li/Ge alloy formation seen in Figures 37D, 37E was the kinetically preferred phase in place of reduced P. Specifically, Ge0 and LixGey together with Li3PS4 and Li7PS6 were the most possible decompositions based on the valences from XPS. Note that at an external pressure of 3MPa and hence reduced Keff on the interfaces, both Ge and P reductions were observed even at a high rate of 2 mA/cm2 (Fig.37F), consistent with the general trend predicted at low Keff in Fig.36B. However, the P reduction might still be kinetically rate- limited, as the most reduced state of Li3P, as predicted in Fig.36B at Keff = 0 GPa and observed in Fig.37C from interface chemical reaction, was not observed.
These two competing reactions with thermodynamic and kinetic preferences, respectively, can be understood by considering a current dependent overpotential for each of these two competing
Figure imgf000078_0005
reactions
Figure imgf000078_0001
. This
Figure imgf000078_0002
term would arise from kinetic effects such as ohmic losses, etc. When current is smal
Figure imgf000078_0003
disappears, thus the thermodynamic overpotential (?) dominates and favors the ground state decomposition products of Figure 36. However, at high currents, ?? begins to dominate and favors those metastable phases, such as LixGey at high
Figure imgf000078_0004
in our computations, which are not shown in Fig.36 as those are all ground state phases in each voltage range.
The impedance profiles before and after CV test (Fig.45A) under 100 MPa or 3 MPa were compared in Fig.45B and 45C after fitting with the model shown in Fig.45D. The calculated Rbulk (bulk resistance) and Rct (charge transfer resistance, here was majorly interface resistance) are listed in Table 5. The Rct (38.8Ω) under 100 MPa is much smaller than that under 3 MPa (395.4Ω) due to a better contact at high pressure. After CV test, there is hardly any change of Rbulk for the battery under 100 MPa, while that of battery under 3 MPa increases from 300Ω to 600Ω. The significantly elevated resistance was attributed to more severe decomposition of LGPS under ineffective mechanical constriction. Again, from electrochemical test, it is proven that the degree of decomposition is significantly inhibited under optimum constriction conditions.
Table 5. Calculated Rbulk and Rct
Figure imgf000079_0001
Conclusion
A lithium-graphite composite allows the application of a high external pressure during the test of solid- state batteries with LGPS as electrolyte. This creates a high mechanical constriction on the materials level that contributes to an excellent rate performance of Li/G-LGPS-G/Li symmetric battery. After cycling at high current densities up to 10 mA cm-2 for such solid-state batteries, cycling can still be performed normally at low rates, suggesting that there is no lithium dendrite penetration or short circuit. The reduction pathway of LGPS decomposition under different mechanical constrictions are analyzed by using both experimental XPS measurements and DFT computational simulations. It shows, for the first time, that under proper mechanical constraint, the LGPS reduction follows a different pathway. This pathway, however, can be influenced kinetically by the high current density induced overpotential. Therefore, the decomposition of LGPS is a function of both mechanical constriction and current density. From battery cycling performance and impedance test, it is shown that high mechanical constriction along with the kinetically limited decomposition pathway reduces the total impedance and realizes a LGPS-lithium metal battery with excellent rate capability.
Methods
Electrochemistry
Graphite thin film is made by mixing active materials with PTFE. The weight ratio of graphite film is graphite: PTFE = 95:5. All the batteries are assembled using a homemade pressurized cell in an argon-filled glovebox with oxygen and water <0.1 ppm. The symmetric battery (Li/G-LGPS-G/Li or Li- LGPS-Li) was made by cold pressing three layers of Li(/graphite)-LGPS powder- (graphite/)Li together and keep at different pressures during battery tests. The batteries were charged and discharged at different current densities with the total capacity of 0.25mAh cm-2 for each cycle. A LiCoO2 half battery was made by cold pressing Li/graphite composite-LGPS powder-Cathode film using a hydraulic press and keep the pressure at 100-250 MPa. The LiCoO2 were coated with LiNbO3 using sol-gel method. The weight ratio of all the cathode films was active materials: LGPS: PTFE = 68:29:3. Battery cycling data were obtained on a LAND battery testing system. The cyclic
performance was tested at 0.1 C at 25℃. The CV test (Li/G-LGPS-LGPS/C) was conducted on a Solartron 1400 cell test system between OCV to 0.1V with the scan rate of 0.1 mV/s. The LGPS cathode film for CV test is made with LGPS: super P: PTFE = 87:10:3.
Material characterization
XRD: The XRD sample was prepared by hand milling LGPS powder with lithium metal and/or graphite with weight ratio = 1:1 in a glovebox. The powder mixtures were put on a hotplate and heated to the nominal temperature (500 °C) for 36 hours and then characterized by XRD. XRD data were obtained using a Rigaku Miniflex 6G. The mixtures of LGPS and graphite before and after high temperature treatment were sealed with Kapton film in an argon-filled glovebox to prevent air contamination.
SEM and XPS: Cross-section imaging of the pellet of Li/graphite-LGPS-graphite-Li was obtained by a Supra 55 SEM. The pellet was broken into small pieces and attached onto the side of screw nut with carbon tape to make it perpendicular to the beam. The screw nuts with samples were mounted onto a standard SEM stub and sealed into two plastic bags inside an argon-filled glove box. FIB-SEM imaging was conducted on an FEIHelios 660 dual-beam system. The XPS was obtained from a Thermo Scientific K-Alpha+. The samples were mounted onto a standard XPS sample holder and sealed with plastic bags as well. All samples were transferred into vacuum environment in about 10 seconds. All XPS results are fitted through peak-differentiating and imitating via Avantage.
Computational Methods All DFT calculations were performed using the Vienna Ab-initio Simulation Package (VASP) following the Material Project calculation parameters.32A K-point density of 1000 kppa, a cutoff of 520 eV, and the VASP recommended pseudopotentials were used. Mechanically constrained phase diagrams were calculated using Lagrange minimization schemes as outlined in Ref.13 for effective moduli of 0, 5, 10 and 15 GPa. All Li-Ge-P-S phases in the Material Project database were considered. Bader charge analysis and spin polarized calculations were used to determine charge valence.
Example 5 - In this work, we focused on how the external application of either high-pressure or isovolumetric conditions can be used to stabilize LGPS at the materials level through the control at the cell-level. This advances beyond the microstructural level mechanical constraints present in previous works, where particle coatings were used to induce metastability. Under proper mechanical conditions, we show that the stability window of LGPS can be widened up to the tool testing upper limit of 9.8 V. Synchrotron X- ray diffraction (XRD) and x-ray absorption spectroscopy (XAS) that measure the structure changes of LGPS before and after high-voltage holding show, for the first time, direct evidence of LGPS straining during these electrochemical processes. Both thermodynamic and kinetic factors are further considered by comparing density functional theory (DFT) simulations and x-ray photoelectron spectroscopy (XPS) measurements for decomposition analysis beyond the voltage stability window. These results suggest that mechanically-induced metastability stabilizes the LGPS up to approximately 4V. Additionally, from 4-10V, the local stresses experienced by decomposition amid rigid mechanical constraints leads to kinetic stability. Combined, mechanically-induced metastability and kinetic stability allow expansion of the voltage window from 2.1V to nearly 10V. To demonstrate the utility of this approach for practical battery systems, we construct fully solid-state cells using this method with various cathodes materials. Li4Ti5O12 (LTO) anodes are paired with LiCo0.5Mn1.5O4 (LCMO), LiNi0.5Mn1.5O4 (LNMO) and LiCoO2 (LCO) cathodes to demonstrate the high-voltage stability of constrained LGPS. To further probe the electrochemical window of LGPS, we report the first all-solid-state battery based on lithium metal and LiCo0.5Mn1.5O4, which can be charged to 6-9 V and cycled up to 5.5 V. Results
To illustrate how mechanical constraint influences the electrochemical stability of LGPS, cyclic voltammetry (CV) tests of LGPS+C/LGPS/Li cells were performed (Figure 46A). Three batteries were pre-pressed with 1, 3, or 6 tons (T) of force (78 MPa, 233 MPa and 467 MPa, respectively) in the assembly and then tested in normal Swagelok batteries. The external pressure of a tightened Swagelok battery was calibrated as a few MPa, giving a quasi-isobaric battery testing condition. In addition, one battery was initially pressed at 6T and then fastened in a homemade pressurized cell with a constantly applied external pressure calibrated as about 200 MPa during the battery test, enforcing a quasi- isovolumetric battery testing environment. The density of the LGPS pellets after being pre-pressed at 1, 3, and 6T were 62%, 69% and 81%, respectively, of the theoretical density of single crystal LGPS. The morphology of LGPS pellets after pressing is shown in Fig. 51A. The density of pellet in the pressurized cell calculated from an in-situ force-displacement measurement (Figure 51B), however, was already close to 100% beyond 30 MPa external pressure.
As shown in Figure 46A, in Cyclic Voltammetry (CV) test there exists a threshold voltage beyond which each cell begins to severely decompose. These thresholds were 4.5 V, 5V and 5.8V for those isobaric cells pre-pressed at 1T, 3T and 6T, respectively. The isovolumetric cell, however, was charged up to 9.8V and showed no obvious decomposition. In the low-voltage region (Figure. 46B), two minor decomposition peaks can be seen at ~3 V and ~3.6 V for the isobaric cells, where decreasing peak intensity was observed at increasing pressure in the pre-press step. On the contrary, the isovolumetric cell completely avoids these peaks. The in-situ resistance of batteries in these four cells were measured by impedance spectroscopy at different voltages during the CV tests (Figure 46C). Higher pressure in pre-press here was found to improve the contact among particles and thus reduce the initial resistance in solid-state battery systems (at 3V in Fig. 46C). However, when the CV test was conducted toward high voltages, the resistance increased much faster in the isobaric cells, indicating that the LGPS in cathode undergoes certain decomposition in the condition of weak mechanical constriction. In contrast, there was almost no change of resistance for the battery tested using the isovolumetric cell. It is worth noting that the voltage stability window of crystalline LGPS toward high voltage was expanded from 2.1 V to around 4.0 V by mechanical constriction induced metastability, the stabilities of 5V to 10V observed in the batteries in Fig.46A far beyond 4 V suggest a different phenomenon.
The synchrotron XRD of LGPS from the isovolumetric cell, as shown in Figure 46D, indicates the general crystal structure of LGPS after CV test up to 9.8 V remains unchanged. However, the broadening of XRD peaks was observed after high-voltage CV scan at 7.5V and 10V (Figures 46E and 52). The peak broadening with increasing 2q angles (Figure 46F) was found to follow the strain broadening mechanism rather than the size broadening. Note that no obvious strain broadening was observed at 3.2V.
This strain effect was further elucidated from XAS measurement and analysis. Figure 46G shows the P and S XAS peaks of pristine LGPS compared with the ones after CV scan up to 3.2V and 9.8V in liquid or solid-state batteries. In the conditions of no mechanical constraint (denoted as 3.2V-L), where LGPS and carbon were mixed with binder and tested in a liquid battery, both P and S show obvious peak shift toward high energy and the shape change, indicating significant global oxidation reaction and rearrangement of local atomic environment in LGPS in the liquid cell. Whereas the P and S peaks don’t show any sign of global oxidation in solid state batteries, as no peak shift is observed. However, it is worth noting that the shoulder intensity increases at 2470 eV and 2149 eV in P and S spectra, respectively. An ab initio multiple scattering simulation of P XAS in LGPS with various strain applied to the unit cell is shown in Figures 46H. A comparison between experiment and simulation suggests that the increase of shoulder intensity in XAS here might be caused by the negative strain, i.e., the compression experienced by crystalline LGPS after CV scan and holding at high voltage. If we connect the strain broadening in XRD with the shoulder intensity increase in XAS, and simultaneously considering that no obvious decomposition current was observed in the CV test up to 10V, a physical picture emerges related to the small local decomposition under proper mechanical constriction. Under a constant external pressure around 150 MPa with nearly zero porosity in the LGPS pellet, macroscopic voltage decomposition of LGPS was largely inhibited kinetically beyond the voltage stability window, i.e. 4.0 V, giving no global transfer of Li+ ion and electron, and hence no decomposition current in CV test. However, small local decomposition inside and between LGPS particle was still able to form. Since decomposition in LGPS is with positive reaction strain, such small local decomposition will exert a compression to the neighboring crystalline LGPS under a mechanically constrictive environment, inducing the strain broadening observed in XRD and the shoulder intensity increase observed in XAS. The fact that both XRD and XAS are ex situ measurements supports our picture on the materials level that such local decomposition induced local strain, once formed, won’t be easily released due to kinetic barriers, even after the external pressure on the battery cell level has been removed. Namely, proper mechanical conditions can lead to a mechanically-induced metastability in LGPS from 4.0V to 10V without obvious decomposition current in the CV test. Our results here provide direct evidences that the electrochemical window of ceramic sulfides can be significantly widened by the proper application of mechanical constraints. In theory, given an unconstrained reaction in which LGPS decomposes with a Gibbs energy change of
Figure imgf000083_0009
the reaction can be inhibited by the application of a mechanical constraint with effective bulk modulus
Figure imgf000083_0008
Figure imgf000083_0002
Where s the reference state volume and ???? is the stress-free reaction dilation– in other words
Figure imgf000083_0014
is the fractional volume change of LGPS following decomposition in the absence of any applied stress. The effective bulk modulus of equation one is the bulk modulus of the ceramic sulfide
Figure imgf000083_0012
added in parallel with the mechanical constraint as given in equation 28:
Figure imgf000083_0001
Minimization of free energy in the mechanically constrained ensemble allows for calculating the expanded voltage window and the ground state decomposition products. Using ab-initio data, Figure 47A shows the results of such calculations for LGPS at four levels of mechanical constraint
Figure imgf000083_0013
0, 5, 10, 15^???) in the voltage range of 0-10V. Figure 47A1 shows the energy above the hull, or the magnitude of the decomposition energy. An energy above the hull of 0 eV atom-1 indicates that thermodynamically the LGPS is the ground state product, whereas an elevated value indicates that the LGPS will decay. The region in which the energy above the hull is nearly zero ( < 50 meV for thermal tolerance) is seen to increase in upper voltage limit from approximately 2.1V to nearly 4V. Figure 47A2 shows the ground state pressure corresponding to the free energy minimization. The pressure is given by where ? corresponds to the fraction volume transformation of LGPS to the products that
Figure imgf000083_0003
Figure imgf000083_0004
minimize the free energy. The ground state pressure reaches 4GPa in the high voltage limit at ???? = 15^???, corresponding well to the level of local strain used in the XAS simulation of strained LGPS in Fig. 46H. Figure 47A3 shows the total specific lithium capacity of the ground state products, which predicts that LGPS electrolyte will not provide more lithium capacity, or make further decomposition, beyond 5V under any Keff below 15 GPa.
The exact decomposition products predicted by DFT without considering the thermal tolerance are shown in Figure 47B in the entire voltage range at different with the exact reaction equations listed
Figure imgf000083_0011
in Table 7. This simulation actually predicts thermodynamically how the small local decomposition reaction induced by electrochemical driving force, as discussed in Fig.46, quantitively changes under mechanical constrictions. The elemental valence states in the decomposition can thus be directly compared with the XPS measurement that is sensitive to the chemical valence information on the particle surface (Fig. 47C, D), providing complementary information to the bulk sensitive XAS. Stoichiometric LGPS is comprised of valence states
Figure imgf000083_0010
As LGPS undergoes the formation of lithium metal (Li1+ ^ Li0) at high voltages, remaining elements must become oxidized. For
Figure imgf000083_0005
, our simulation in Figure 47B suggests that sulfur is the most likely to be oxidized, forming
Figure imgf000083_0006
above 2.3V and S0 (elemental sulfur) above 3.76V. From the DFT simulation of Bader charge,
Figure imgf000083_0007
shows very similar charge state, and obviously higher than S2- in LGPS, which is consistent with the large amount of oxidized S observed in XPS for LGPS in the liquid cell after CV scan to 3.2V and hold for 10 hours (Fig.47C2). Similarly, the oxidization of P in the same 3.2V liquid cell is observed to form P5+ in PS43- (Fig.47D2). This suggests that the thermodynamically favored decomposition is in fact representative of the decomposition that occurs experimentally in the liquid cell with ???? = 0 (as opposed to an alternative kinetically favored decomposition under mechanical constriction).
In contrast, the calculated thermodynamic stability limit of LGPS reaches nearly 4V at ???? ^= ^15^???. Accordingly, there was no oxidization of S and a very small amount of oxidized P was observed in the condition of strongly constrained LGPS at 3.2V in Figures 47C3 and D3. This small amount of oxidized P could be attributed to the ineffective constraint from the device or the voltage is close to the thermodynamic voltage. Furthermore, beyond the voltage stability limit for the case of 9.8 V, the solid- state battery showed less oxidized S or P than it was expected. Note that from Figure 47B, there is supposed to be the decomposition of LGPS into S element and oxidized P in Li7PS6 or Li2PS3. However, this thermodynamic pathway was bypassed. Beyond this thermodynamic stability, there is kinetical factor to stabilize sulfide electrolyte under high mechanical constraint.
The application of the mechanical constraint can greatly reduce the speed at which ceramic sulfides decay as depicted in Figure 53. Upon sufficient slowing of the decay rate, the effective stability– the “mechanically-induced kinetic stability”– was sufficiently high as to allow battery operation. For example, if the electrolyte only decays one part per million per charge cycle, then it was sufficiently stable for practical battery designs that only need last thousands of cycles.
The proposed mechanism for mechanically-induced kinetic stability is depicted in Figure 53. Within a given particle of LGPS that is undergoing decomposition, the particle can be partitioned into three regions. The first two are the decomposed and pristine regions, which are indicated in Figure 53 (top) by the mole fraction of decomposed LGPS (?? = 1 for purely decomposed, ?? = 0 for pristine). The third region is the interface, where the mole fraction transitions from 0 to 1. The propagation direction of the decomposition front is controlled by thermodynamic relation of Equation 1. If Equation 1 is satisfied, the front will propagate inwards, preferring the pristine LGPS. Accordingly, the LGPS will not decompose. When Equation 1 is violated, the front will propagate into the LGPS and ultimately consume the particle.
However, even when Equation 1 is violated, the speed with which the front propagates into the pristine LGPS will still be influenced by the application of mechanical constraint. This is illustrated in Figure 53 (bottom). As the decomposition front propagates, there must exist ionic currents tangential to the front’s curvature. This requires the presence of an overpotential to accommodate the finite conductivity of the front for each elemental species. The ohmic portion of the overpotential is given by the sum of equation 3, where is the resistivity of the front for each species ? at the pressure (?) that is present at the front, is the characteristic length scale of the decomposed morphology, and ?? is the ionic current density.
Figure imgf000084_0001
Given that ????? can quickly grow with constriction, it is to be expected that this overpotential becomes significant at high pressures. This effect can be seen by comparing the expected constriction with prior molecular dynamics results of constricted cells. The pressure on the decomposition front is given by ? = ???????? and the elastic volume strain of the material at that pressure is ? Since the
Figure imgf000085_0003
strain of a single lattice vector is approximately the strain of the ab-plane of LGPS near the
Figure imgf000085_0004
front is expected to be on the order of For well constrained systems where ???? »
Figure imgf000085_0002
?????????, this strain can easily reach 4%, exceeds 30% at high voltages. Given that the
Figure imgf000085_0005
activation energy for Li migration in LGPS is predicted to increase from 230 meV to 590 meV upon constriction by 4%, the rate at which lithium reordering can occur decreases by a factor of:
Figure imgf000085_0001
This many order of magnitude reduction in the possible reordering rate can explain why, for any voltage below 10V, the isovolumetric cell showed virtually no decomposition current.
Figure 48 shows the galvanostatic cycling along with their cyclability performance of all-solid-state batteries, using LCO, LNMO and LCMO as cathode, LGPS as a separator and LTO as anode. The battery tests were performed in the pressurized cell, where the cells were initially pressed with 6T then fastened in bolted [quasi]-isovolumetric cell. It should be noted that LCO is the most common and widely used cathode material, included in commercial Li-ion batteries, with a plateau at approximately 4 V against Li+/Li, whereas LNMO is considered one of the most promising high voltage cathode materials with a flat operating voltage at 4.7 V versus Li+/Li. The high rate test of LCO full battery is shown in Figure 55. The charge and discharge curves of LCO and LNMO are depicted in Figs.48A1 and 48B1, respectively. Both batteries show a flat working plateau centered at 2 V (3.5 V vs Li+/Li) for LCO and 2.9 V (4.4 V vs. Li+/Li) for LNMO in the first discharge cycle. Moreover, both of them exhibit excellent cyclability performance, as can be observed in Figs.48A2 and B2, with a capacity fading of just 9% in the first 360 cycles for LCO and 18% in the first 100 cycles for LNMO. This is an indication that the decomposition or interfacial reaction of the cathode materials with LGPS was not very severe. These results are in good agreement with the CV tests reported in Figure 46, where it was shown that mechanical constraint can inhibit the decomposition of LGPS and widen its operational voltage range to much higher values than those previously reported. Moreover, to further probe the stability of LGPS, previously synthesized LCMO was chosen as cathode due to the fact that it presents even a higher operating working plateau than LNMO. Figure. 48A3 depicts the battery test curves of LCMO versus LTO. In both charge and discharge profiles, two plateaus can be observed centered at approximately 2.2 V and 3.2 V (3.7 V and 4.7 V versus Li+/Li) in the discharge curve of the first cycle, which are associated to the oxidation reactions of Mn3+/Mn4+ and Co3+/Co4+, respectively. As it is shown in Figure 48B3, upon cycling some capacity fading was observed, which may be attributed to the side reactions between LCMO and LGPS at high voltage state and corresponds to an 33% in the 50th cycle. Therefore, in contrast to previously reported results, which claims that the stability window of LGPS was limited to a low voltage range, here we show that LGPS can be used as the electrolyte material in high-voltage- cathode all-solid-state batteries, showing a relatively good cycling performance even when the charging plateau is as high as 3.8 V (5.3 V versus Li+/Li). Figures 48C1-48D3 show the XPS measured binding energy of electrons in LGPS before and after battery cycles using LCO, LNMO and LCMO as cathodes. Each element can become oxidized either by chemical reaction with the cathode material (chemical oxidation) or the delithiation of the LGPS by the application of a voltage (electrochemical oxidation). As decipted in Figures 48C1-48D3, those electrons in the characteristic region of sulfur bonded electrons show a peak shift towards a higher energy state after cycling, indicating that the sulfur has become electrochemically oxidized. The presence of oxidized sulfur in the pristine samples is indiciative of the degree of chemical reaction with the cathode material.
XAS measurement shows a pre-edge on the intensity of S element while no pre-edge is found from P (Figures 48E and 56), given that S, instead P, is bonded with trasition metal, no matter from coating materials or cathode materials. Althought the interface reaction is evaliated by the mechanical constraint, there is still a ceterin amount of side reactions happens from the direct contract between cathode materials and LGPS. More interface reactions occur after battery cycles.
Interfacial reactions between two materials (i.e. LGPS and a cathode material) present computational challenges as ab-initio simulations of the interface present unique burdens. Instead, the preferred method to simulate both chemical and electrochemical stabilities of interfaces are the so-called pseudo-phase (also known as pseudo-binary) methods. In these methods, a linear combination of the materials of interest are taken and represented as a single phase with both composition and energy given by the linear combination. This phase is the pseudo-phase. Conventional stability calculations can then be applied to the pseudo-phase to estimate the reaction energy of the interface. Figures 49A-D and Table 6 give the results for chemical reaction pseudo-phase calculations for LGPS + LNO, LCO, LNMO, and LCMO. In Figures 49A-D, the atomic fraction of the cathode material (or LNO) is swept from 0 to 1 (representing pure LGPS to pure cathode or LNO). Whichever value of atomic fraction makes the reaction energy the most negative represents the worst-case reaction and is termed ??. Table 6 gives these ?? values for each interface, along with the worst-case reaction energy, the decomposed products, and an additional pseudo-phase that represents the decomposed interface. This pseudo-phase that represents the decomposed interface, also known as the interphase, can be used to calculate how the decomposed interface will further decay as the battery is cycled. Figure 49E-G show the electrochemical stability of the LGPS+LNO interphase. Note that the chemical reaction between LGPS and the cathode material happens as soon as the materials come in contact during cathode film assembly. This is in contrast with the electrochemical reactions which do not occur until the external circuit assembly is attached. Thus, a major difference between the two is that chemical reactions occur before pressurization/cell assembly whereas the electrochemical reactions occur afterwards. Since the chemical reactions occur in the absence of a fully assembled cell, the initial reactions always occur at ???? = ? (the electrochemical reactions occur at the ???? of the completed assembly). Table 6. Chemical reaction data for the interface between LGPS and either LNO, LCO, LCMO, or LNMO. ???? is the worst-case reaction energy between the two phases and ?? is the atomic fraction of the non-LGPS phase that is consumed in this worst-case scenario.‘Products’ lists the phases that result from this worst-case reaction.‘Chemical decomp pseudo-phase’ is the application of pseudo- phase theory to the set of products in‘products.’ It represents an artificial phase with a linear combination of composition, energy, and volume of its constituent phases.
Figure imgf000087_0001
Figures 49B-D show that the chemical reaction energies for LCO, LNMO, and LCMO are 345, 322, and 335 meV atom-1, respectively. Despite being coated with LNO, which has a much lower reaction energy of 124 meV atom-1 (Figure 49A), the coating is not perfect allowing some contact with LGPS which results in the chemical oxidation of sulfur seen in the pristine samples of Figures 48C-48E. Figures 49E-G show that the products that result from the chemical reaction of LGPS and LNO (which constitute the LGPS-LNO interphase) also experience mechanically-induced metastability. Thus, in a full cell in which the cathode particles are coated with LNO, proper constriction (such as those batteries depicted in Figure 48) should lead to mechanically-induced metastability both within the bulk of the solid-electrolyte as well as at the interface with the cathode materials. As a general rule, LGPS interfaces were more likely to experience mechanically-induced metastabilities with insulators (such as LNO) than with conductors (such as LCO, LNMO, and LCMO). The reason for this is that when the interphase oxidizes to form lithium metal, the lithium metal will form locally if the interface is between two electronically insulating materials. If one of the two phases is conducting, however, the lithium ions can migrate to the anode and thus form a non-local phase. In the latter case, the local reaction dilation will be greatly reduced as the volume of the formed lithium phase will not be included in the local volume change. In contrast, if the lithium metal phase forms locally, it contributes to a larger local volume change and, hence, a larger reaction dilation. For this reason, coating cathode materials in an insulator such as LNO is needed in order for constraints to lead mechanically-induced metastability on the interface of the LGPS.
Usually, lithium metal is soft and which leads to the difficulty of applying pressure due to the immediate short of lithium through the bulk solid electrolyte. In order to probe the high voltage capability of pressurized LGPS in the system of lithium metal solid-state battery, lithium metal was used as anode with a graphite layer as a protection layer, which allows high pressure applied during battery test. Firstly, lithium metal-LCO batteries were made at different mechanical conditions using Swagelok, aluminum pressurized cell and stainless-steel pressurized cell, as shown in Figure 57. Again, the interface reaction and decomposition reaction in the strongest constraint condition is the lowest. A similar structure was applied to make a higher-voltage lithium metal battery using LCMO as cathode, where the cell was initially pressed with 6T. It is shown in Figure. 58 that graphite protection layer alleviate the interface reaction between lithium metal and LGPS. As shown in Figure 59, The decomposition of LGPS itself is very small in the condition of strong mechanical constraint, it contributes very small decomposition current as shown in Figure 59. As depicted in Figure.50A, the LCMO cathode then can be charged up to 9 V, which simulates the high-voltage charge status of not-yet-discovered high-voltage redox chemistries. Discharging capacities of 99, 120, 146, 111 mAh/g are obtained by charging LCMO at 6,7,8,9 V, respectively (Figure 50A). This indicates that the extra lithium capacity comes from the LCMO’s higher voltage state. Although there are more side reactions after the battery is charged to voltages above 8 V, the battery is seen to maintain the capability of cycling even up to 9V. This high- voltage cycling demonstrates the high electrochemical window of over 9 V for constrained LGPS. At highly delithiated state, cathode materials usually show poor electrochemical stability and the reaction between cathode materials and electrolyte is also more severe.
To contrast this performance with conventional electrolytes, Figure 50B depicts organic liquid electrolyte failing at nearly 5V. However, the solid-state battery tested under isovolumetric conditions can be charged up to 9 V (Fig.50A) without evidence of a decomposition plateau. Moreover, a battery cycling at 5.5 V and tested under isovolumetric conditions (initially pressed with 6T) (Figure 50C), shows a stable cycling performance and high Columbic efficiency even at high cut-off voltage of 5.5 V, in contrast to the liquid battery (Figure 50B). Although the performance of lithium metal-LCMO battery is not as good as full battery due to the mechanical softness of lithium metal, this result still shows that, unlike liquid electrolytes, solid-state electrolytes are a better platform to run high-voltage cathode materials. In summary, we demonstrate how mechanical constraint widens the stability of ceramic solid electrolyte, pushing up its electrochemical window to levels beyond organic liquid electrolytes. A CV test shows that properly designed solid-state electrolytes working under isovolumetric conditions can operate up to nearly 10 V, without clear evidence of decomposition. A mechanism for this mechanically induced kinetic stability of sulfides solid-electrolytes is proposed. Moreover, based on this understanding, it has been shown how several high-voltage solid-state battery cells, using some of the most commonly used and promising cathode materials, can operate up to 9 V under isovolumetric conditions. Therefore, the development of high-voltage solid-state cells is not compromised by the stability of the electrolyte anymore. We anticipate that this work is an import breakthrough for the development of new energy storage systems and cathode materials focused on very-high voltage (>6V) electrochemistry. Method Sample characterization Structural Analysis
Routine XRD data were collected in a Rigaku Miniflex 6G diffractometer working at 45 kV and 40 mA, using CuKa radiation (wavelength of 1.54056 Å). The working conditions were 2q scanning between 10–80 º, with a 0.02 º step and a scan speed of 0.24 seconds per step.
Electrochemical characterization
The LGPS+C/LGPS part of the cells were pellets which were made by pressing the powder at 1T, 3T, 6T, respectively, and put into Swagelok or the homemade pressurized cell. In the CV test, voltage starting from the open circuit voltage to 10 V was ramped, during which the decomposition currents at each voltage were measured. The CV test was conducted on a Solartron 1400 electrochemical test system between OCV to 3.2V, 7.5V, and 9.8V, respectively, with the scan rate of 0.1 mV/s. The CV scan was followed by a voltage hold for 10 hours to make sure the decomposition is fully developed, and it was scanned back to 2.5V before any other characterizations. The electrochemical impedance spectroscopy (EIS) was conducted on the same machine in the range of 3 MHz to 0.1 Hz.
For all-solid-state batteries, the electrode and electrolyte layers were made by a dry method which employs Polytetrafluoroethylene (PTFE) as a binder and allows to obtain films with a typical thickness of 100-200 µm. Additionally, two different kinds of all-solid-state batteries were assembled, using Li4Ti5O12 (LTO) or lithium (Li) metal as anode. In any case, the composite cathode was prepared by mixing the active materials (LiCo0.5Mn1.5O4, LiNi0.5Mn1.5O4 or LiCoO2) and Li10GeP2S12 (LGPS) powder in a weight ratio of 70:30 and 3% extra of PTFE. This mixture was then rolled into a thin film. On the one hand, for those all-solid-state batteries which use LTO as anode, a separator of LGPS and PTFE film was employed with a weight ratio of 95:5. The anode composition consists in a mixture of LGPS, LTO and carbon black in weight ratio 60:30:10 and 3% extra of PTFE. Finally, the Swagelok battery cell of cathode film (using LiCo0.5Mn1.5O4, LiNi0.5Mn1.5O4 or LiCoO2 as active material) /LGPS film/LTO film was then assembled in an argon-filled glove box. The specific capacity was calculated based on the amount of LTO (30 wt%) in the anode film. The galvanostatic battery cycling test was performed on an ArbinBT2000 work station at room temperature. On the other hand, when lithium metal was used as anode, a Li metal foil with a diameter and thickness of ½” and 40 µm, respectively, was connected to the current collector. In order to prevent interface side reactions, the Li foil was covered by a 5/32” diameter carbon black film with a weight ratio of carbon black and PTFE of 96:4. After loading the negative electrode into a Swagelok battery cell, 70 mg of pure LGPS powder, which acts as a separator, was added and slightly pressed. Finally, ~1 mg film of the cathode composite LCMO was inserted and pressed up to 6 Tn (0.46 GPa) to form the battery, which final configuration was LCMO/LGPS pellet/graphite film+Li metal. For high voltage test in Figure 50A, the battery is charged to 0.3C followed by 30 mins rest and discharged at 0.1C. All batteries in Figure 50 are test at high temperature of 55℃. Computational Simulation
All ab-initio calculations and phase data were obtained following the Material Project calculation guidelines in the Vienna Ab-initio Software Package (VASP). The mechanically-induced metastability calculations were performed following the LaGrangian optimization methods outlined in Small 1901470, 1–14 (2019) and J. Mater. Chem. A (2019). doi:10.1039/C9TA05248H). Pseudo-phase calculations were performed following the methods of J. Mater. Chem. A 4, 3253–3266 (2016), Chem. Mater. 28, 266–273 (2016), and Chem. Mater.29, 7475–7482 (2017). Other embodiments are in the claims.

Claims

What is claimed is: CLAIMS
1. A rechargeable battery, comprising a first electrode, a second electrode, and a solid state electrolyte disposed therebetween, wherein the solid state electrolyte comprises a sulfide comprising an alkali metal, wherein the solid state electrolyte is under a volumetric constraint sufficient to stabilize the solid state electrolyte during electrochemical cycling.
2. The rechargeable battery of claim 1, wherein the volumetric constraint exerts a pressure between about 70 and about 1,000 MPa on the solid state electrolyte.
3. The rechargeable battery of claim 1, wherein the volumetric constraint exerts a pressure between about 100 and about 250 MPa on the solid state electrolyte.
4. The rechargeable battery of claim 1, wherein the volumetric constraint provides a voltage stability window of between 1 and 10 V.
5. The rechargeable battery of claim 1, wherein the solid state electrolyte has a core shell morphology.
6. The rechargeable battery of claim 1, where the alkali metal is Li, Na, K, Rb, or Cs.
7. The rechargeable battery of claim 1, wherein the solid state electrolyte comprises SiPS, GePS, SnPS, PSI, or PS.
8. The rechargeable battery of claim 1, wherein the solid state electrolyte is Li10SiP2S12, Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3.
9. The rechargeable battery of claim 1, wherein the first electrode is the cathode and comprises LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or LiCo0.5Mn1.5O4.
10. The rechargeable battery of claim 1, wherein the second electrode is anode and comprises lithium metal, lithiated graphite, or Li4Ti5O12.
11. The rechargeable battery of claim 1, wherein the volumetric constraint provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa.
12. A rechargeable battery comprising a first electrode, a second electrode, and a solid state electrolyte disposed therebetween, wherein the second electrode is an anode comprising an alkali metal and graphite.
13. The rechargeable battery of claim 12, wherein the battery is under a pressure of about 70-1000 MPa.
14. The rechargeable battery of claim 13, wherein the battery is under a pressure of about 100-250 MPa.
15. The rechargeable battery of claim 12, wherein the alkali metal and graphite form a composite.
16. The rechargeable battery of claim 12, where the alkali metal is Li, Na, K, Rb, or Cs.
17. The rechargeable battery of claim 12, wherein the solid state electrolyte comprises SiPS, GePS, SnPS, PSI, or PS.
18. The rechargeable battery of claim 12, wherein the solid state electrolyte is Li10SiP2S12,
Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3.
19. The rechargeable battery of claim 12, wherein the first electrode is the cathode and comprises LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or LiCo0.5Mn1.5O4.
20. The rechargeable battery of claim 12, wherein the battery is under an external stress that provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa.
21. A rechargeable battery comprising a first electrode, a second electrode, and a solid state electrolyte disposed therebetween, wherein the solid state electrolyte comprises a sulfide comprising an alkali metal; and the battery is under isovolumetric constraint.
22. The rechargeable battery of claim 21, wherein the isovolumetric constraint is provided by compressing the solid state electrolyte under a pressure of about 3-1000 MPa.
23. The rechargeable battery of claim 21, where the alkali metal is Li, Na, K, Rb, or Cs.
24. The rechargeable battery of claim 21, wherein the solid state electrolyte comprises SiPS, GePS, SnPS, PSI, or PS.
25. The rechargeable battery of claim 21, wherein the solid state electrolyte is Li10SiP2S12,
Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3.
26. The rechargeable battery of claim 21, wherein the first electrode is the cathode and comprises LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or LiCo0.5Mn1.5O4.
27. The rechargeable battery of claim 12, wherein the isovolumetric constraint provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa.
28. A rechargeable battery, comprising a first electrode, a second electrode, and a solid state electrolyte disposed therebetween, wherein:
a) the solid state electrolyte comprises a sulfide comprising an alkali metal; and
b) at least one of the first or second electrodes comprises an interfacially stabilizing coating material.
29. The rechargeable battery of claim 28, wherein the first electrode is the cathode and comprises a material selected from Table 1.
30. The rechargeable battery of claim 28, wherein the coating material of the first electrode comprises a material selected from Table 2.
31. The rechargeable battery of claim 28, where the alkali metal is Li, Na, K, Rb, or Cs.
32. The rechargeable battery of claim 28, wherein the solid state electrolyte comprises SiPS, GePS, SnPS, PSI, or PS.
33. The rechargeable battery of claim 28, wherein the solid state electrolyte is Li10SiP2S12,
Li10GeP2S12, or Li9.54Si1.74P1.44S11.7Cl0.3.
34. The rechargeable battery of claim 28, wherein the first electrode is the cathode and comprises LiCoO2, LiNi0.5Mn1.5O4, V Li2CoPO4F, LiNiPO4, Li2Ni(PO4)F, LiMnF4, LiFeF4, or LiCo0.5Mn1.5O4.
35. The rechargeable battery of claim 28, wherein the battery is under an external stress that provides a mechanical constriction on the solid state electrolyte between about 1 to about 100 GPa.
36. The rechargeable battery of claim 28, wherein the battery is under a pressure of about 70-1000 MPa.
37. The rechargeable battery of claim 36, wherein the battery is under a pressure of about 100-250 MPa.
38. A method of storing energy comprising applying a voltage across the first and second electrodes and charging the rechargeable battery of any one of claims 1-37.
39. A method of providing energy comprising connecting a load to the first and second electrodes and allowing the rechargeable battery of any one of claims 1-37 to discharge.
PCT/US2019/063354 2018-11-26 2019-11-26 Solid state batteries WO2020112843A1 (en)

Priority Applications (7)

Application Number Priority Date Filing Date Title
CN201980090173.8A CN113454825A (en) 2018-11-26 2019-11-26 Solid-state battery
US17/297,228 US20210408580A1 (en) 2018-11-26 2019-11-26 Solid state batteries
CA3120864A CA3120864A1 (en) 2018-11-26 2019-11-26 Solid state batteries
AU2019387113A AU2019387113A1 (en) 2018-11-26 2019-11-26 Solid state batteries
JP2021529271A JP2022509633A (en) 2018-11-26 2019-11-26 Solid state battery
KR1020217019956A KR20210100651A (en) 2018-11-26 2019-11-26 solid state battery
EP19890968.1A EP3888175A4 (en) 2018-11-26 2019-11-26 Solid state batteries

Applications Claiming Priority (2)

Application Number Priority Date Filing Date Title
US201862771319P 2018-11-26 2018-11-26
US62/771,319 2018-11-26

Publications (2)

Publication Number Publication Date
WO2020112843A1 true WO2020112843A1 (en) 2020-06-04
WO2020112843A8 WO2020112843A8 (en) 2020-07-23

Family

ID=70853668

Family Applications (1)

Application Number Title Priority Date Filing Date
PCT/US2019/063354 WO2020112843A1 (en) 2018-11-26 2019-11-26 Solid state batteries

Country Status (8)

Country Link
US (1) US20210408580A1 (en)
EP (1) EP3888175A4 (en)
JP (1) JP2022509633A (en)
KR (1) KR20210100651A (en)
CN (1) CN113454825A (en)
AU (1) AU2019387113A1 (en)
CA (1) CA3120864A1 (en)
WO (1) WO2020112843A1 (en)

Cited By (1)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
WO2022094412A1 (en) * 2020-10-30 2022-05-05 President And Fellows Of Harvard College Batteries with solid state electrolyte multilayers

Families Citing this family (4)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
WO2019104181A1 (en) 2017-11-22 2019-05-31 President And Fellows Of Harvard College Solid state electrolytes and methods of production thereof
CN112701345B (en) * 2020-12-29 2022-04-12 长三角物理研究中心有限公司 Super-hydrophobic material capable of conducting lithium ions as well as preparation method and application thereof
CN114958351B (en) * 2022-06-22 2023-09-15 旭宇光电(深圳)股份有限公司 Ultraviolet excited blue-violet fluorescent powder, preparation method thereof and light-emitting device
WO2024096107A1 (en) * 2022-11-04 2024-05-10 住友化学株式会社 Battery

Citations (4)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US20140193691A1 (en) * 2011-05-02 2014-07-10 Tomoyoshi Ueki Nonaqueous electrolyte secondary battery
US20160276641A1 (en) * 2015-03-19 2016-09-22 Toyota Jidosha Kabushiki Kaisha Nonaqueous electrolyte secondary battery
WO2018098494A1 (en) * 2016-11-28 2018-05-31 Quantumscape Corporation Pressurized electrochemical cell
WO2019104181A1 (en) * 2017-11-22 2019-05-31 President And Fellows Of Harvard College Solid state electrolytes and methods of production thereof

Family Cites Families (8)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
JP2010034002A (en) * 2008-07-31 2010-02-12 Idemitsu Kosan Co Ltd Lithium battery and lithium battery mounting device
JP2014035987A (en) * 2012-08-10 2014-02-24 Toyota Motor Corp METHOD FOR MANUFACTURING Si-CONTAINING ACTIVE MATERIAL LAYER, METHOD FOR MANUFACTURING SOLID-STATE BATTERY, Si-CONTAINING ACTIVE MATERIAL LAYER, AND SOLID-STATE BATTERY
EP2997611A4 (en) * 2013-05-15 2016-12-14 Quantumscape Corp Solid state catholyte or electrolyte for battery
JP6123642B2 (en) * 2013-11-08 2017-05-10 トヨタ自動車株式会社 All-solid battery charging system
CN107851840B (en) * 2015-06-01 2022-08-23 气动覆层科技有限责任公司 Nano-engineered coatings for solid state electrolytes
WO2018056082A1 (en) * 2016-09-21 2018-03-29 株式会社村田製作所 Solid electrolyte and all-solid battery
EP3483971A4 (en) * 2016-12-26 2019-11-13 Showa Denko K.K. All-solid-state lithium ion battery
KR20180094184A (en) * 2017-02-13 2018-08-23 전자부품연구원 All-solid battery using sulfide-based solid electrolyte and method of manufacturing the same

Patent Citations (4)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US20140193691A1 (en) * 2011-05-02 2014-07-10 Tomoyoshi Ueki Nonaqueous electrolyte secondary battery
US20160276641A1 (en) * 2015-03-19 2016-09-22 Toyota Jidosha Kabushiki Kaisha Nonaqueous electrolyte secondary battery
WO2018098494A1 (en) * 2016-11-28 2018-05-31 Quantumscape Corporation Pressurized electrochemical cell
WO2019104181A1 (en) * 2017-11-22 2019-05-31 President And Fellows Of Harvard College Solid state electrolytes and methods of production thereof

Non-Patent Citations (2)

* Cited by examiner, † Cited by third party
Title
See also references of EP3888175A4 *
WU, F ET AL.: "Advanced sulfide solid electrolyte by core-shell structural design", NATURE COMMUNICATIONS 9, vol. 9, no. 4037, 2 October 2018 (2018-10-02), XP055615492 *

Cited By (1)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
WO2022094412A1 (en) * 2020-10-30 2022-05-05 President And Fellows Of Harvard College Batteries with solid state electrolyte multilayers

Also Published As

Publication number Publication date
US20210408580A1 (en) 2021-12-30
JP2022509633A (en) 2022-01-21
CN113454825A (en) 2021-09-28
EP3888175A4 (en) 2022-11-16
KR20210100651A (en) 2021-08-17
CA3120864A1 (en) 2020-06-04
AU2019387113A1 (en) 2021-06-17
WO2020112843A8 (en) 2020-07-23
EP3888175A1 (en) 2021-10-06

Similar Documents

Publication Publication Date Title
Banerjee et al. Interfaces and interphases in all-solid-state batteries with inorganic solid electrolytes
Zuo et al. A mechanistic investigation of the Li10GeP2S12| LiNi1-x-yCoxMnyO2 interface stability in all-solid-state lithium batteries
WO2020112843A1 (en) Solid state batteries
Nanda et al. Lithium degradation in lithium–sulfur batteries: insights into inventory depletion and interphasial evolution with cycling
Doerrer et al. High energy density single-crystal NMC/Li6PS5Cl cathodes for all-solid-state lithium-metal batteries
Dong et al. Oxide cathodes: functions, instabilities, self healing, and degradation mitigations
Wachs et al. Online monitoring of transition-metal dissolution from a high-Ni-content cathode material
Tian et al. Electron and ion transfer across interfaces of the NASICON-type LATP solid electrolyte with electrodes in all-solid-state batteries: a density functional theory study via an explicit interface model
Yang et al. Computational auxiliary for the progress of sodium-ion solid-state electrolytes
Davis et al. Rate limitations in composite solid-state battery electrodes: revealing heterogeneity with operando microscopy
Qin et al. Revisiting the electrochemical lithiation mechanism of aluminum and the role of Li‐rich phases (Li1+ xAl) on capacity fading
Sångeland et al. Probing the interfacial chemistry of solid-state lithium batteries
Kim et al. Thermally driven interfacial degradation between Li7La3Zr2O12 electrolyte and LiNi0. 6Mn0. 2Co0. 2O2 cathode
Barai et al. Investigation of delamination-induced performance decay at the cathode/LLZO interface
Song et al. Metastable Li1+ δMn2O4 (0≤ δ≤ 1) spinel phases revealed by in operando neutron diffraction and first-principles calculations
Hänsel et al. Favorable interfacial chemomechanics enables stable cycling of high-Li-content Li–In/Sn anodes in sulfide electrolyte-based solid-state batteries
Gavilán-Arriazu et al. Kinetic Monte Carlo simulations applied to Li-ion and post Li-ion batteries: a key link in the multi-scale chain
Ghigna et al. Operando x-ray absorption spectroscopy on battery materials: a review of recent developments
WO2019104181A1 (en) Solid state electrolytes and methods of production thereof
Morgan et al. From Atoms to Cells: Multiscale Modeling of LiNi x Mn y Co z O2 Cathodes for Li-Ion Batteries
Zabara et al. Operando investigations of the interfacial electrochemical kinetics of metallic lithium anodes via temperature-dependent electrochemical impedance spectroscopy
Li et al. Highly stable and encapsulation-microstructural cathode derived by self-pressurization behavior in Na-halides-based all-solid-state batteries
Jena et al. In Situ/Operando Methods of Characterizing All-Solid-State Li-Ion Batteries: Understanding Li-Ion Transport during Cycle
Riegger et al. Instability of the Li7SiPS8 solid electrolyte at the lithium metal anode and interphase formation
Malaki et al. Advanced analytical characterization of interface degradation in Ni-rich NCM cathode co-sintered with LATP solid electrolyte

Legal Events

Date Code Title Description
121 Ep: the epo has been informed by wipo that ep was designated in this application

Ref document number: 19890968

Country of ref document: EP

Kind code of ref document: A1

ENP Entry into the national phase

Ref document number: 3120864

Country of ref document: CA

ENP Entry into the national phase

Ref document number: 2021529271

Country of ref document: JP

Kind code of ref document: A

NENP Non-entry into the national phase

Ref country code: DE

ENP Entry into the national phase

Ref document number: 2019387113

Country of ref document: AU

Date of ref document: 20191126

Kind code of ref document: A

ENP Entry into the national phase

Ref document number: 20217019956

Country of ref document: KR

Kind code of ref document: A

ENP Entry into the national phase

Ref document number: 2019890968

Country of ref document: EP

Effective date: 20210628