ACS Publications. Most Trusted. Most Cited. Most Read
Fabrication of Oxide-Based All-Solid-State Batteries by a Sintering Process Based on Function Sharing of Solid Electrolytes
My Activity

Figure 1Loading Img
  • Open Access
Energy, Environmental, and Catalysis Applications

Fabrication of Oxide-Based All-Solid-State Batteries by a Sintering Process Based on Function Sharing of Solid Electrolytes
Click to copy article linkArticle link copied!

  • Miyuki Sakakura
    Miyuki Sakakura
    Department of Materials Design Innovation Engineering, Graduate School of Engineering, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8603, Japan
  • Kazutaka Mitsuishi
    Kazutaka Mitsuishi
    Research Center for Advanced Measurement and Characterization, National Institute for Materials Science, 1-2-1 Sengen, Tsukuba 305-0047, Japan
  • Toyoki Okumura
    Toyoki Okumura
    Research Institute of Electrochemical Energy, National Institute of Advanced Industrial Science and Technology (AIST), 1-8-31 Midorigaoka, Ikeda, Osaka 563-8577, Japan
  • Norikazu Ishigaki
    Norikazu Ishigaki
    Department of Materials Design Innovation Engineering, Graduate School of Engineering, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8603, Japan
  • Yasutoshi Iriyama*
    Yasutoshi Iriyama
    Department of Materials Design Innovation Engineering, Graduate School of Engineering, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8603, Japan
    *Email: [email protected]
Open PDFSupporting Information (1)

ACS Applied Materials & Interfaces

Cite this: ACS Appl. Mater. Interfaces 2022, 14, 43, 48547–48557
Click to copy citationCitation copied!
https://doi.org/10.1021/acsami.2c10853
Published October 3, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

Garnet-type Li7La3Zr2O12 (LLZ) has advantages of stability with Li metal and high Li+ ionic conductivity, achieving 1 × 10–3 S cm–1, but it is prone to react with electrode active materials during the sintering process. LISICON-type Li3.5Ge0.5V0.5O4 (LGVO) has the advantage of less reactivity with the electrode active material during the sintering process, but its ionic conductivity is on the order of 10–5 S cm–1. In this study, these two solid electrolytes are combined as a multilayer solid electrolyte sheet, where 2 μm thick LGVO films are coated on LLZ sheets to utilize the advantages of these two solid electrolytes. These two solid electrolytes adhere well through Ge diffusion without significant interfacial resistance. The LLZ–LGVO multilayer is combined with a LiCoO2 positive electrode and a lithium metal anode through annealing at 700 °C. The resultant all-solid-state battery can undergo repeated charge–discharge reactions for over 100 cycles at 25 or 60 °C. The LGVO coating suppresses the increases in the resistance from the solid electrolyte and interfacial resistance induced by annealing by ca. 1/40. As with sulfide-based all-solid-state batteries, function sharing of solid electrolytes will be a promising method for developing advanced oxide-based all-solid-state batteries through a sintering process.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2022 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Oxide-based all-solid-state batteries (Ox-SSBs) are attracting attention as candidates for next-generation rechargeable batteries that can realize high energy density while maintaining safety and reliability. Thin-film-type Ox-SSBs with a few mA h sizes were commercialized approximately 20 years ago, where positive electrode, solid electrolyte, and negative electrode materials were sequentially piled up by vapor deposition processes such as RF sputtering and vacuum evaporation. (1,2) In recent years, multilayer-type Ox-SSBs have been commercialized by applying multilayer ceramic capacitor technology, where both electrodes and the solid electrolyte are cosintered, (3,4) and some of them obtain capacities over 20 mA h. (4) Because most oxide-based solid electrolytes, especially those with high Li+ conductivity, are hard and fragile crystalline ceramics, such a sintering process is one of the practical ways to develop advanced Ox-SSBs.
A severe problem accompanying the preparation of Ox-SSBs through the sintering process is the side reactions between the electrode and solid electrolyte. These side reactions provide a large interfacial resistance in many cases. (5−7) Among the various oxide-based solid electrolytes, garnet-structured Li7La3Zr2O12 (LLZ) has been examined through experimental and theoretical approaches because of its high ionic conductivity close to 1 × 10–3 S cm–1 at room temperature in addition to its stability against Li metal. (8−11) However, LLZ highly reacts with many electrode active materials during the sintering process. For example, when a LiCoO2 (LCO) cathode is sintered with LLZ at 700 °C, La2CoO4 is formed as a byproduct at the interface, and the interfacial resistance seriously increases. (5)
An effective way to compensate for the disadvantages of a solid electrolyte is to combine the solid electrolyte with another solid electrolyte. Such function sharing of solid electrolytes has been commonly examined in sulfide-based SSBs from the viewpoint of suppressing side reactions in charge–discharge reactions, (12,13) reducing space charge layer formation, (14,15) or extending the potential window of solid electrolytes. (16) In the case of Ox-SSBs, NASICON-type Li1+xAlxTi2–x(PO4)3 (LATP), which is unstable with Li metal, (17,18) is combined with lithium phosphorus oxynitride glass electrolyte (LiPON), which is stable with Li metal by forming a multilayer-mosaic solid electrolyte interphase (SEI), (19) and then, Li plating–stripping reactions are realized on LiPON/LATP multilayers. (20,21) The concept of function sharing of solid electrolytes will be extended to reduce the reactivity of the electrode/solid electrolyte interface during the sintering process, that is, combining LLZ with another solid electrolyte with less reactivity with electrode materials during the sintering process. Several reports apply Li3BO3 (LBO) with a lower melting point of approximately 700 °C (22,23) as the combined solid electrolyte. Ohta et al. fabricated a LCO–LBO composite layer on a sintered Nb-doped LLZ substrate by a screen printing method and then annealed it at 700 °C for 1 h. (22) The composite electrode ca. 10 μm in thickness can undergo repeated stable charge–discharge reactions. The application of LBO also has advantages in realizing a well-adhered electrode-solid electrolyte interface and then developing bulk-type Ox-SSBs using powder electrode materials. (22,23) However, the ionic conductivity of LBO has been reported to be 2 × 10–6 S cm–1 at room temperature. (22) This ionic conductivity is improved to 5 × 10–6 S cm–1 by Li2+xC1–xBxO3, (24−26) but the ionic conductivity is still 2 orders of magnitude smaller than that of LLZ. Replacing LBO with another material with higher ionic conductivity is a better approach to reduce the Ohmic resistance of solid electrolytes. Okumura et al. developed bulk-type Ox-SSBs by spark plasma sintering, combining LISICON-type Li3.5Ge0.5V0.5O4 (LGVO) with LiNi1/3Co1/3Mn1/3O2 (NCM111). (27) LGVO is stable with various LiCoO2-type electrodes, including NCM111, under sintering at 700 °C, and the resultant Ox-SSBs can undergo repeated stable charge–discharge reactions. The ionic conductivity of LGVO is close to 10–4 S cm–1 at room temperature, (27) which is 1 order of magnitude higher ionic conductivity than that of LBO but smaller than that of LLZ. Therefore, placing a thin layer of LGVO as a protective layer against side reactions on LLZ, in addition to reducing the total Ohmic resistance of the resultant function sharing solid electrolyte, will be useful.
In this work, we prepared LGVO-coated LLZ substrates and investigated whether this kind of function sharing solid electrolyte is effective in suppressing the interfacial reactivity under sintering and then providing a low-resistivity interface. Here, a crystalline LGVO thin film with ca. 2 μm thickness was prepared on an LLZ substrate by aerosol deposition (AD), and the LLZ–LGVO multilayer was annealed at different temperatures. Here, AD is a room-temperature ceramic coating technology, and it has been applied to prepare various coating materials, such as Al2O3, ZrO2, LCO, and electrode-solid electrolyte composite layers. (6,28−31) Then, a LCO thin film was prepared on the LGVO side of the LLZ–LGVO multilayer using AD and annealed at 700 °C. Finally, the resultant LLZ–LGVO/LCO was combined with a lithium metal anode, and an Ox-SSB, Li/LLZ/LGVO/LCO, was prepared. We will show that the LLZ–LGVO multilayer provides a well-adhered low-resistivity interface through Ge diffusion by annealing at 700–900 °C. The 2 μm LGVO coatings effectively reduce the resistance from the solid electrolyte and interfacial resistance by ca. 1/40 compared to the noncoated Ox-SSB. An Ox-SSB using the LLZ–LGVO multilayer can undergo repeated stable charge–discharge reactions for over 100 cycles.

Experimental Section

Click to copy section linkSection link copied!

Preparation of LGVO Powder for AD

LGVO powder for AD was synthesized from Li4GeO4 and Li4VO4 as the starting material. (27) The Li4GeO4 powder was prepared by mixing LiOH–H2O powder (>99.0%, Nakarai Tesuque) and GeO2 powder (>99.99%, Nakarai Tesque) in a stoichiometric molar ratio and then fired at 600 °C for 12 h in O2. Then, the mixture was pressed into pellets at a pressure of 360 MPa and sintered at 700 °C for 12 h in O2. The Li4VO4 powder was prepared by mixing LiOH–H2O powder (>99.0%, Nacalai Tesque) and V2O5 powder (>99.0%, Kanto Chemical) in a stoichiometric molar ratio and then fired at 450 °C for 12 h in O2. Then, the mixture was pressed into pellets at a pressure of 360 MPa and sintered at 550 °C for 12 h in O2. The resultant Li4GeO4 and Li3VO4 powders were mixed in an agate mortar at a molar ratio of 1:1, then formed into pellets at a pressure of 360 MPa and sintered at 700 °C for 12 h in O2 to obtain LGVO pellets. Electrochemical impedance spectroscopy (EIS) measurements were carried out at 25 °C by applying an alternating current (AC) amplitude of 10 mV (500 kHz to 100 mHz) to check the ionic conductivity of pelletized LGVO. The LGVO pellets were ground in a planetary ball mill (P-5, Fritsch) at a rotational speed of 100 rpm for 2 h in an Ar atmosphere. After removing large particles through a 100 μm mesh sieve, starting powders for AD were prepared. Field emission scanning electron microscopy (FE-SEM) (SU8030, Hitachi Ltd.) and X-ray diffraction (XRD) (40 kV, 40 mA, Cu Kα, Rigaku Ltd.) measurements were carried out to investigate the morphology and crystallinity of the powders, respectively.

Preparation of LGVO Films by AD and Characterization

Ta-doped sintered LLZ substrates (Li6.6La3Zr1.6Ta0.4O12, 0.5 mm in thickness, 10 × 10 mm, Toshima) were used as base substrates. These substrates were first polished by SiC sandpaper (#400), and then, surface residual impurities such as Li2CO3 and polishing dust were removed by acid etching (immersion in 1 mol dm–3 HCl dissolved in H2O, washing with H2O, and then drying under vacuum). (32) The acid-etched LLZ substrates were set in a vacuum evacuated AD chamber, and they were fixed to a substrate holder so that the deposition area was a rectangular shape of 8 × 10 mm (deposition area: 0.8 cm2). The starting powder amount for AD was 2 g. The powders were stored in the AD vessel maintained at 150 °C during deposition. The carrier gas of Ar flowed into the vessel at 15 L min–1, and the resultant aerosol was sprayed onto the substrate through a nozzle. During AD, the substrate was moved back and forth without heating. After AD, substrates were transferred into an Ar-filled glovebox, and the surface dust generated during AD was wiped off using a blower and wiping paper. The resultant LLZ–LGVO multilayers were annealed at 700, 900, or 1050 °C for 2 h under an Ar atmosphere. These annealed multilayers are denoted ML-700, ML-900, and ML-1050, respectively, and as-prepared multilayers without annealing are denoted ML-AP. Both surface and cross-section textures were observed by FE-SEM, and element distributions were analyzed by energy-dispersive X-ray spectroscopy (EDX) (X-max, Horiba Ltd.) at an acceleration voltage of 15 kV. Crystallinity variations of the MLs with annealing were evaluated using XRD. The ionic conductivities of both the MLs and the polished LLZ substrate were measured by EIS. The measurement cells were blocking electrode cells, where Au thin films (3 mm in diameter) were deposited on both sides of these substrates by RF sputtering after polishing the LLZ surface with SiC polishing paper (#400). Cross-sectional TEM specimens were prepared using a focused ion beam (JIB-4501, JEOL Ltd). Their structure and elemental compositions were examined using scanning transmission electron microscopy (STEM; JEM-ARM200F, JEOL Ltd.) equipped with EDX.

Fabrication and Electrochemical Measurements of Ox-SSBs

Crystalline LCO thin films (ca. 1 μm in thickness) were deposited on the LGVO side of ML-900 by AD using laboratory-made LCO particles (D50: 2.2 μm). (33) The AD preparation parameters were the same as those for the LGVO film except for the deposition area (6 mm in diameter). The obtained samples were annealed at 700 or 900 °C for 2 h in an Ar atmosphere. The reactivity of the LCO formed on LGVO with annealing was preliminarily investigated by Raman spectroscopy (NRS-5100, Japan Spectroscopy) and STEM equipped with EDX, and an appropriate annealing temperature was determined to be 700 °C. After annealing at 700 °C, a Au current collector film was deposited on the LCO film by RF sputtering in a 7 mm square region. A Li metal anode film was formed on the LLZ side of ML-900 by vacuum evaporation with an area 7 mm in diameter after polishing the LLZ surface with a SiC polishing paper (#400) in an Ar-filled glovebox (dew point < −80 °C). The resultant Ox-SSB (Li/LLZ/LGVO/LCO/Au) was sealed in a coin-type cell (CR2032, Hosen Corporation) and then heated at 165 °C to obtain a well-adhered Li/LLZ interface. Additionally, Ox-SSBs without a LGVO coating layer were prepared without and with annealing at 700 °C in the same manner to clarify the role of the LGVO coating layer.
Charge–discharge measurements of the Ox-SSBs with ML-900 were performed in the constant-current constant-voltage mode at 25 or 60 °C between 3.0 and 4.2 V. The initial 25 cycles were performed at 60 °C, in which the charge current density was fixed at 2.9 μA cm–2 and the discharge current density was varied from 2.9 to 286 μA cm–2. Then, 11 cycles were performed at 25 °C, in which the charge current density was fixed at 2.9 μA cm–2 and the discharge current density was varied from 2.9 to 57 μA cm–2. Finally, the charge current density and operating temperature were fixed at 5 μA cm–2 and 60 °C, respectively, and the charge–discharge reactions were repeated for 75 cycles. The discharge current density was varied from 5 to 500 μA cm–2 in the first 25 cycles and fixed at 50 μA cm–2 in the last 50 cycles. During these charge–discharge reactions, EIS measurements were carried out at 3.0 and 4.0 V after a given charge–discharge reaction by applying an AC amplitude of 10 mV within the frequency range of 200 kHz to 100 mHz. In the case of the Ox-SSBs without LGVO coating, charge–discharge reactions were repeated for 10 cycles at 5 μA cm–2, where the initial 5 cycles were measured at 60 °C and the later 5 cycles were conducted at 25 °C. Additionally, EIS measurements were carried out after the 5th cycle (60 °C) and the 10th cycle (25 °C) at 3.0 and 4.0 V.

Results and Discussion

Click to copy section linkSection link copied!

Preparation of LLZ–LGVO MLs

Figure 1a shows an FE-SEM image of the synthesized LGVO powder for AD, and the inset shows a magnified image of the powder. The LGVO powder formed as secondary particles less than 10 μm in diameter, consisting of submicron-sized primary particles. Figure 1b shows the XRD pattern of the LGVO powder. The diffraction pattern from the LGVO powder (black spectrum) was in good agreement with the reported pattern (red spectrum (34)), and impurity peaks were not detected (ND), indicating that single-phase LGVO was synthesized. Figure 1c shows the EIS spectrum of a sintered LGVO pellet (density: 72%) measured at 25 °C. Au films were deposited on both sides of the LGVO pellet (Au/LGVO/Au). The inset spectrum is a magnified image of a higher frequency region. The spectrum consists of one semicircular arc to ca. 8000 Ω, followed by a straight line to the real axis in the lower frequency region. Considering that this semicircular arc is assigned to the ion-conduction resistance of LGVO, the ionic conductivity was calculated to be 5 × 10–5 S cm–1 at 25 °C. This value was comparable to reported values. (27,35)

Figure 1

Figure 1. (a) FE-SEM image of LGVO particles for AD. The inset shows a magnified SEM image of a particle. (b) XRD patterns of the LGVO particles (black) and reported reference data (red). (c) Nyquist plots of a Au/LGVO/Au symmetrical cell at 25 °C. The inset displays a magnified image of a higher frequency region. The numbers in these plots indicate data points measured at 10n Hz.

Figure 2 summarizes surface and cross-sectional FE-SEM images of the MLs (ML-AP, ML-700, ML-900, and ML-1050). Figure 2a–d shows surface SEM images of the MLs. The inset in each surface SEM image is an optical image of each ML, where a LGVO film was formed in an 8 × 10 mm region on a 10 mm square LLZ substrate. The LGVO film in ML-AP (Figure 2a) appears to be composed of fine particles, and this situation was the same in ML-700 (Figure 2b). In contrast, grain growth was visibly observed in ML-900 (Figure 2c) and more clearly in ML-1050 (Figure 2d). Figure 2e–h shows cross-sectional SEM images of the MLs. For each sample, LGVO films ca. 2 μm in thickness uniformly form on the LLZ substrates. Grain growth proceeds with the increase in annealing temperature to over 900 °C, as shown in Figure 2c,d, but the LGVO film does not peel off from the LLZ substrate. The LLZ region of both ML-AP and ML-700 is porous and densified in ML-900 and ML-1050. As mentioned in the experimental section, the LLZ substrates used in these MLs were acid-etched to remove surface impurities. This acid etching provided a porous LLZ region around the substrate, but such a porous region appeared to be densified with sintering over 900 °C. This densification is caused by the sintering of LLZ itself (Figure S1), and Ge diffusion from LGVO, as described below, may also assist it. (36) Figure S2 summarizes the SEM and EDX mapping images of Ge Lα, V Kα, and La Lα for ML-AP (Figure S2a–d), ML-700 (Figure S2e–h), ML-900 (Figure S2i–l), and ML-1050 (Figure S2m–p). The averaged Ge/V ratio (N3) of ML-AP is 0.99, which is almost consistent with the ideal ratio (1.00). In contrast, MLs with annealing over 700 °C exhibit decreased average Ge/V ratio to 0.79–0.90, indicating that LGVO reacts with LLZ with Ge diffusion. This reactive sintering will work as a driving force for generating a well-adhered interface, which will reasonably explain the in-plane growth of LGVO on LLZ with annealing. It has been reported that Ge can be doped into LLZ and occupy Li or La sites depending on the doping amount. (36) The appropriate amount of Ge doping of less than 1 wt % forms a solid solution with a cubic phase in addition to improving the ionic conductivity, while an excess amount of Ge doping of over 1 wt % provides tetragonal-phase LLZ, detected as (024) peak splitting, in addition to the appearance of impurity diffraction peaks. (36) The XRD patterns from both ML-700 and ML-900 do not show such peak splitting or impurity peaks, as shown in Figure 4. Therefore, the amount of Ge diffusion inside LLZ is expected to be in the solid solution range (less than 1 wt %). Both Ge Lα and V Kα signals are observed in most of the areas on the MLs, but some areas are shown as a dark color, where the La Lα signal from the LLZ substrate appears strong. These results indicate that LGVO films are formed almost uniformly on the LLZ substrates on a microscale, but they may not be coated or may be coated in a thinner state on part of the MLs on the macroscale.

Figure 2

Figure 2. (a–d) Surface and (e–f) cross-sectional FE-SEM images of (a,e) ML-AP, (b,f) ML-700, (c,g) ML-900, and (d,h) ML-1050. The insets in (a–d) display optical images of the MLs.

Figure 3a,b show cross-sectional TEM images of ML-AP and ML-900. The LGVO films in ML-AP are formed by small-sized grains with voids, and the LGVO/LLZ interface remains a disconnected region along the interface. In contrast, the LGVO films in ML-900 are well sintered and densely adhere to LLZ. Figure 3c,d shows the EDX profile measured inside the LLZ of ML-AP and ML-900, and the analyzed values are summarized in each table, where both the mass % and at. % were estimated by the ζ-factor method. (37) EDX analysis could ND Ge in the LLZ region of ML-AP (Figure 3c). In contrast, Ge was detected at 0.09 at. % in ML-900 (Figure 3d) and widely distributed from the interface, although the detected value was in the lower detection limit range and should be considered a reference value. When Ge occupies the Li site, (36) the wide-range distribution of Ge may be explained by Ge diffusion through the Li+ conduction pathway. This wide-range diffusion, not condensed around the interface, is consistent with the reduction in the Ge/V ratio induced by annealing in the EDX measurements. V appears to be present in the LLZ region from the EDX image, but this is because the V Kα signal overlaps with the La Lα and O Kα signals. The amount of V was estimated by the ζ-factor method, (37) but V was ND in either ML-AP or ML-900, suggesting that V will not diffuse toward LLZ.

Figure 3

Figure 3. Cross-sectional TEM images of (a) ML-AP and (b) ML-900. EDX measurements around the LLZ/LGVO interface of (c) ML-AP and (d) ML-900. The EDX signals were from O Kα, V Kα, Ge Kα, Zr Lα, La Lα, and Ta Lα. The tables in Figure 2c,d summarize the mass %, σ, and at. % evaluated in the LLZ region in the red frame.

Figure 4 shows the XRD patterns of the MLs. The diffraction patterns of the LLZ substrate after polishing by SiC polishing paper (#400) and LGVO powder are also shown as references. ML-AP presents a broad diffraction pattern from the LGVO film in addition to diffraction peaks from the LLZ substrate. The broad diffraction pattern from the LGVO film is attributed to the crystallinity degradation by AD. In the cases of ML-700 and ML-900, the diffraction peaks from the LGVO film appear sharp, indicating that the crystallinity of the LGVO film has recovered. Although Ge diffusion inside LLZ was suggested by EDX analysis, no impurity peaks were detected. The diffused amount of Ge (0.09 at. %) inside LLZ is estimated to be 0.13 wt % GeO2 against the Li7La3Zr2O12 composition, which is in the range of a solid solution of Ge (less than 1 wt %). (36) Additionally, the Ge/V ratio of 0.79–0.90 is within the range of the solid solution of LGVO, (35) and then, a Li3+xGexV1–xO4 (x = 0.48–0.44) film will be formed on LLZ. The diffraction peak intensity from the LGVO film further increases in ML-1050, and the crystallinity of the LGVO film improves; however, diffraction peaks assigned to La2Zr2O7 newly appear at 28.45, 47.44, and 56.19°. (38) La2Zr2O7 will be generated by thermal decomposition of LLZ, probably through lithium loss due to high-temperature annealing at 1050 °C. To summarize these XRD data, ML-AP provides a lower-crystalline LGVO layer, while ML-1050 provides highly crystalline LGVO, but the resistive impurity phase of La2Zr2O12 is generated as a byproduct. In the case of ML-700 and ML-900, LGVO is highly crystalline, and there are no impurity XRD peaks. Thus, both ML-700 and ML-900 are suitable for structural evaluation.

Figure 4

Figure 4. XRD patterns of ML-AP (blue), ML-700 (orange), ML-900 (red), and ML-1050 (purple). Additionally, XRD patterns of the LLZ substrate after polishing with SiC sandpaper (black) and LGVO powder (green) are shown as references. Diffraction peaks marked by * are assigned to La2Zr2O7.

Figure 5a shows the EIS spectra of the blocking electrode cell of the MLs and the polished LLZ substrate. Because these substrates have slightly different thicknesses due to the polishing treatment of the LLZ substrates, all the data are normalized by the thickness (Ω cm–1). The EIS spectra exhibit characteristic spectra for blocking electrodes and diverge in lower frequency regions. Figure 5b–f show magnified images of higher frequency regions of LLZ and ML-AP, -700, -900, and -1050. One semicircular arc is observed in the high-frequency range except for ML-AP (Figure 5c). The total ionic conductivity (σtotal) of the MLs was calculated from the diameters of the semicircular arcs (Rtotal), and the values are displayed in each spectrum. The LLZ substrate has a σtotal of 7.3 × 10–4 S cm–1. Assuming that the ionic conductivity of the 2 μm thick LGVO films on LLZ is 5 × 10–5 S cm–1 for all samples and that there is no resistance at the LLZ/LGVO interface, the σtotal of the ML is estimated to be 6.9 × 10–4 S cm–1. The σtotal values of ML-700 and ML-900 are 7.0 × 10–4 and 6.8 × 10–4 S cm–1, respectively. These values are almost consistent with the ionic conductivity estimated from the sum of each resistance of LLZ and LGVO, indicating that there is no significant interfacial resistance at LLZ/LGVO. Although the composition of LGVO may change from the initial Li3.5Ge0.5V0.5O4 with the highest ionic conductivity to Li3+xGexV1–xO4 (x = 0.48–0.44) via Ge diffusion, this compositional change will result in only a slight decrease in the ionic conductivity. (27) In the case of ML-1050, σtotal decreases to 3.3 × 10–4 S cm–1, which is attributed to the formation of resistive La2Zr2O7 as a byproduct detected by XRD. Additionally, ML-AP exhibits quite a large semicircular arc out of the scale in Figure 5c, and σtotal is estimated to be 4.6 × 10–5 S cm–1 from Figure 5a. This decreased ionic conductivity is attributed to the degraded crystallinity of the LGVO film with voids, as shown in Figures 2 and 4, in addition to the formation of the disconnected region along the interface (Figure 3). These results indicate that both ML-700 and ML-900 are appropriate in terms of ionic conductivity.

Figure 5

Figure 5. (a) Nyquist plots of ML-AP (blue), ML-700 (orange), ML-900 (red), ML-1050 (purple), and the LLZ substrate (green) measured at 25 °C. Magnified Nyquist plots in a higher frequency region of the (b) LLZ substrate, (c) ML-AP, (d) ML-700, (e) ML-900, and (f) ML-1050. The numbers in each plot indicate data points measured at 10n Hz. The thickness of each substrate is used to normalize all the resistances.

From these results, both ML-700 and ML-900 are considered appropriate MLs in this study. ML-900 was selected as a model substrate to fabricate Ox-SSBs through a sintering process in the subsequent evaluation.

Electrochemical Properties of the Ox-SSBs Using ML-900

We checked the reactivity of LGVO with LCO under sintering by Raman spectroscopy before fabricating Ox-SSBs using ML-900. Thin films of LCO were formed on LGVO pellets using AD at room temperature, and then, the LCO/LGVO ML was sintered at 700 or 900 °C in Ar. Figure S3a shows a cross-sectional FE-SEM image of a LCO thin film fabricated on a LGVO pellet. A LCO thin film with a thickness of ca. 1 μm was formed on the LGVO pellet. Figure S3b shows Raman spectra of the as-deposited LCO film and the LCO film after annealing at 700 or 900 °C. The as-deposited LCO film exhibits two Raman bands (A1g and Eg) characteristic of layered-structured LCO. (39) These two bands were observed after annealing at 700 °C but disappeared at 900 °C. These results clarify that LGVO-LCO significantly reacts with annealing at less than 900 °C. Figure S4 shows a cross-sectional TEM image and EDX line profile of LCO/LGVO interface annealing at 700 °C. A mutual diffusion layer may be formed around the interface, but its thickness is less than 20 nm. Wide-range mutual diffusion of Co, V, and Ge was ND.
Based on the above results, a thin film of LCO was deposited on the LGVO side of ML-900 using AD and annealed at 700 °C. Then, a lithium metal anode was deposited on the LLZ side of the ML-900/LCO, and Ox-SSBs (Li/ML-900/LCO) were fabricated. Figure S5 shows a cross-sectional SEM image (Figure S5a) and an EDX mapping image (Figure S5b). Both LCO (ca. 1 μm) and LGVO (ca. 2 μm) films formed an ML on an LLZ substrate. Although we could not measure the correct loading amount of the LCO film because the AD conditions in this work slightly etched the substrate, the expected amount was ca. 0.5 mg/cm2 assuming that the LCO thin film has the ideal density of LCO (5.0 g/cm3). The lithium plating amount during the charging process is estimated to be ca. 0.3 μm.
Figure 6 shows the variation in the discharge capacity of an Ox-SSB with repetition of charge–discharge cycles at different current densities. The reaction temperature was swung: 1st–25th cycles at 60 °C, 26th–36th cycles at 25 °C, and 37th–137th cycles at 60 °C. The charge current density was fixed at 2.9 μA cm–2 for the 1st–36th cycles and 5 μA cm–2 after the 37th cycle, and the discharge current density was changed from 2.9 to 500 μA cm–2. The discharge capacity slightly decreased with the cycles, and the discharge capacity after the 137th cycle (I = 50 μA cm–2) maintained 88% of that after the 11th cycle (I = 57 μA cm–2). Additionally, the total interfacial resistance (Rint) measured after the given charge–discharge reactions is plotted in Figure 6, where Rint is the sum of the interfacial resistances at Li/LLZ (DE1-R) and LCO/LGVO (DE2-R), as discussed below. Figure 7a displays the 1st, 6th, 11th, 16th, and 21st charge–discharge curves at 60 °C, where the discharge current density was varied from 2.9 to 286 μA cm–2. The Coulombic efficiency in the initial cycle was 89.7%. Figure 7b displays the charge–discharge curves at the 26th, 31st, and 36th cycles at 25 °C, where the discharge current density was varied from 2.9 to 57 μA cm–2. In these charge–discharge curves, the hexagonal-monoclinic phase transition of LCO was not observed at 4.1–4.2 V in Figure 7a, but it clearly appeared in Figure 7b. This situation agrees with the temperature dependency of the hexagonal-monoclinic phase transition. (40) These results indicate that the LCO film on ML-900 maintains good crystallinity after annealing at 700 °C. Figure 7c,d shows the EIS spectra measured at 3.0 and 4.0 V after the 5th (60 °C) and 25th (25 °C) charge–discharge cycles, respectively. The high-frequency intercept with the real axis does not depend on the potential, and the values are 60 Ω at 60 °C and 230 Ω at 25 °C, which is assigned to the resistance of ML-900 considering the frequency in Figure 5e. In contrast, a new semicircular arc appears at 4.0 V. This semicircular arc is displayed as Rint in Figure 7c,d. Rint is divided into two relaxation processes from the distribution of relaxation times (DRT) analysis (Figure S6), which is the sum of the interfacial resistances at Li/LLZ (DE1) and LGVO/LCO (DE2). (32) EIS spectra recorded at 3.0 and 4.0 V after the given charge–discharge cycles are displayed in Figure S7. Table 1 summarizes the DRT-analyzed resistances of the solid electrolyte (R1), Li/LLZ interface (DE1-R), and LGVO/LCO interface (DE2-R) in addition to other values [relaxation times (DE-T), depression factors (DE-P), and error % for the fitting] of these EIS spectra at 4.0 V. Additionally, the equivalent circuit model is displayed in Table 1. Here, the variations in Rint in Figure 6 are the sum of DE1-R and DE2-R. DE1-R slightly increases from 13.2 Ω (after the 5th cycle) to 19.2 Ω (after the 136th cycle) with cycling, which is attributed to the formation of a partially disconnected Li/LLZ interface region. (32) DE2-R increases ca. two times from 46.9 Ω (after the 5th cycle) to 109 Ω (after the 136th cycle). As shown in Figure 6, the discharge capacity slightly decreases with the cycles. This increased interfacial resistance at LGVO/LCO will contribute to the degradation and rise of Rint with cycling. A plausible mechanism for the increasing DE2-R is the formation of a disconnected LGVO/LCO interface originating from the volume change of LCO at 3.0–4.2 V (ca. 2%) during the charge–discharge reactions. (6,26,27,40) An effective way to reduce this degradation will be to replace LCO with other electrode materials with smaller volume changes, such as LiNi1/3Co1/3Mn1/3O2. (27,31) Focusing on the interfacial resistivity of LGVO/LCO at 25 °C measured after the 15th charge–discharge reaction is interesting. DE2-R was estimated to be 514 Ω; thus, the interfacial resistivity was calculated to be 145 Ω cm2. This is almost the same value as LCO/LLZ (150 Ω cm2) interface-modified by 10 nm Nb metal, (41) where the interfacial resistance sensitively varied with the thickness of the Nb coating layer. This means that a low-resistivity LGVO/LCO interface is formed even after annealing at 700 °C.

Figure 6

Figure 6. Variation in the discharge capacity and total interfacial resistance (Rint) of an Ox-SSB (Li/ML-900/LCO) with the charge–discharge reactions measured at 60 and 25 °C. The charge current density was fixed at 2.9 μA cm–2 in the 1st–36th cycles and 5 μA cm–2 in the 37th–137th cycles, and the discharge current density was varied from 2.9 to 286 μA cm–2 during the 1st–36th cycles and 5–500 μA cm–2 during the 37th–91st cycles and fixed at 50 μA cm–2 during the 92nd–137th cycles. The Rint values were measured at 4.0 V after charge–discharge reactions in the 5th, 25th, 36th, 86th, and 136th cycles.

Figure 7

Figure 7. Charge–discharge curves of an Ox-SSB (Li/ML-900/LCO) with 700 °C annealing in the (a) 1st, 6th, 11th, 16th, and 21st cycles and (b) 26th, 31st, and 36th cycles, where the charge current density was fixed at 2.9 μA cm–2 and discharge current density was 2.9 μA cm–2 (1st and 26th), 29 μA cm–2 (6th and 31st), 57 μA cm–2 (11th and 36th), 143 μA cm–2 (16th), and 286 μA cm–2 (21st). Nyquist plots at 3.0 and 4.0 V after the (c) 5th cycle and (d) 25th cycle, where Rint denotes the sum of the interfacial resistances at Li/LLZ and LGVO/LCO. Data were measured at 60 °C in (a,c) and at 25 °C in (b,d). The numbers in (c,d) indicate data points measured at 10n Hz.

Table 1. Variations in EIS Data Analyzed by DRT of an Ox-SSB with ML-900 after the 5th, 36th, 86th, and 136th Charge–Discharge Cycles Shown in Figure 6; R1 (Solid Electrolyte), DE-1 (Li/LLZ Interface), and DE-2 (LCO/LGVO Interface)a
 5th36th86th136th
 valueerror (%)valueerror (%)valueerror (%)valueerror (%)
R1 (Ω)62.13.0 × 10–561.82.9 × 10–564.52.6 × 10–571.62.7 × 10–5
DE1-R (Ω)13.26.9 × 10–413.16.2 × 10–414.84.7 × 10–419.23.4 × 10–4
DE1-T (s)2.82 × 10–63.3 × 10–42.82 × 10–62.8 × 10–42.82 × 10–62.0 × 10–42.82 × 10–61.5 × 10–4
DE1-P0.892.5 × 10–40.902.3 × 10–40.891.9 × 10–40.881.5 × 10–4
DE2-R (Ω)46.91.7 × 10–459.21.2 × 10–476.58.1 × 10–51095.5 × 10–5
DE2-T (s)2.92 × 10–51.9 × 10–43.55 × 10–61.5 × 10–44.75 × 10–51.2 × 10–47.73 × 10–57.5 × 10–5
DE1-P0.784.5 × 10–50.764.0 × 10–50.733.4 × 10–40.7131 × 10–5
a

The equivalent circuit model of the Ox-SSB is shown in the table.

To clarify the role of the LGVO coating layer in ML-900, an all-solid-state battery (Li/LLZ/LCO) without a LGVO coating layer was fabricated by depositing LCO on an acid-treated LLZ substrate by AD and then annealed at 700 °C in the same manner. Figure 8a shows the charge–discharge curves of the 1st–5th cycles at 60 °C. In the 1st cycle, the open-circuit voltage (OCV) was as low as 1.65 V, and a charging capacity of 13 μA h cm–2 was observed under 3.0 V. The Coulombic efficiency over 3.0 V was decreased to 66.6%, with an irreversible capacity of 24 μA h cm–2. In contrast, the discharge capacities after two cycles were almost consistent with those of the SSBs using ML-900 in Figure 7a and delivered 47 μA h cm–2. Here, the hexagonal-monoclinic phase transition of LCO was not observed, as shown in Figure 7a. Figure 8b shows the EIS spectra recorded after the 5th cycle (60 °C) at 3.0 and 4.0 V. The high-frequency intercept with the real axis was potential-independent, and this part was then assigned to the resistance of the solid electrolyte. However, the resistance increased from 60 Ω (Figure 7a) to 2000 Ω (Figure 8a). Additionally, Rint increased from 60 Ω (Figure 7b) to 2500 Ω (Figure 8b). These results indicate that the LGVO coating effectively inhibits the increase in both the solid electrolyte resistance and Rint because of the suppression of side reactions during the sintering process. When charging and discharging reactions were carried out at 25 °C (Figure 8c), the charge–discharge voltage was visibly polarized, and the hexagonal-monoclinic phase transition peaks were not observed, probably because of the inhomogeneous reaction induced by the large interfacial resistance. Additionally, potential-independent high-frequency intercepts were composed of several semicircular arcs; the value increased from 200 to 17000 Ω, and the potential-dependent Rint increased from 650 Ω (Figure 7d) to 25000 Ω (Figure 8d). This increased resistance reasonably explains the large polarization of the charge–discharge curve.

Figure 8

Figure 8. (a) Charge–discharge curves at 60 °C of an Ox-SSB (Li/LLZ/LCO) with 700 °C annealing in the 1st–5th cycles under a fixed charge–discharge current density of 5 μA cm–2. (b) Nyquist plots at 60 °C of a nonannealed Ox-SSB (Li/LLZ) measured at 3.0 V (blue) and 4.0 V (red) after the 5th cycle. (c) Charge–discharge curves at 25 °C of a nonannealed Ox-SSB (Li/LLZ/LCO) in the 6th–10th cycles under a fixed charge–discharge current density of 5 μA cm–2. (d) Nyquist plots at 25 °C of an Ox-SSB (Li/LLZ/LCO) without annealing measured at 3.0 V (blue) and 4.0 V (red) after the 10th cycle. (e) Charge–discharge curves at 60 °C of an Ox-SSB (Li/LLZ/LCO) without annealing in the 1st–5th cycles under a fixed charge–discharge current density of 5 μA cm–2. (f) Nyquist plots at 60 °C of an Ox-SSB (Li/LLZ) without annealing measured at 3.0 V (blue) and 4.0 V (red) after the 5th cycle. (g) Schematic image of the interface properties with LGVO coating and annealing effects.

The use of acid-etched LLZ provides a porous region around the LLZ surface, as mentioned in Figure 2, which may increase the cell resistance due to worse physical bonding. (32) Thus, we prepared an Ox-SSB (Li/LLZ/LCO) without annealing after depositing LCO thin films by AD on acid-treated LLZ, where the maximum heating temperature was 165 °C to obtain a well-adhered Li/LLZ interface. Figure 8e shows the charge–discharge curves at 60 °C of the resultant SSB (Li/LLZ/LCO) without LGVO or any annealing. The OCV increased to 2.51 V, and the capacity below 3.0 V disappeared. The presence of an additional plateau at approximately 3.6 V may be due to the degraded crystallinity in part of the as-prepared AD LCO film. (6) In the first cycle, an irreversible capacity of 26 μA h cm–2 was observed, and the Coulombic efficiency was 67.5%. The discharge capacity after the second cycle was ca. 50 μA h cm–2, comparable to those of the annealed Ox-SSB without LGVO (Figure 8a) and the Ox-SSB using ML-900. In contrast, the Coulombic efficiency in the first cycle was ca. 90% for the Ox-SSB with ML-900 and decreased to 67–68% in Ox-SSBs without LGVO regardless of annealing. The difference in the irreversible capacity in the first cycle may arise from the anodic instability of LLZ, (16,42) and LGVO may have a superior anodic stability to LLZ. When the oxidation of LLZ occurs during the first charging reaction, the interface will be modified to a stabilized state at the voltage, similar to SEI formation on a graphite anode in a lithium-ion battery. After the stabilization of the interface, the later charging reaction will not entail any irreversible capacity under the same charging voltage. Figure 8f shows the EIS spectra recorded after the fifth cycle (60 °C) at 3.0 and 4.0 V. The potential-independent high-frequency intercept with the real axis and Rint were 50 and 70 Ω, respectively. These values were comparable to those of the Ox-SSB with ML-900 (Figure 7c) but 2 orders of magnitude smaller than those of the sintered Ox-SSB without LGVO coating (Figure 8d). These results clarify that the porosity of the LLZ does not influence the drastic increase in the cell resistance and that annealing of LLZ/LCO induces an increase in the cell resistance. In other words, the 2 μm thick LGVO coating effectively suppresses the side reactions at LLZ/LCO and then inhibits the increase in the resistance induced by sintering.
Figure 8g summarizes the properties of the LLZ/LCO interface with LGVO coating and annealing. The LLZ/LCO interface without annealing has low resistivity. Once LLZ/LCO is annealed, the materials react with each other, and a high-resistivity interface is generated. The LGVO layer reacts with LLZ through Ge diffusion induced by annealing, but LLZ/LGVO has low resistivity. In contrast, LGVO/LCO is a low-reactivity and low-resistivity interface, and LLZ/LGVO/LCO provides a totally low-resistivity interface. These results indicate that one strategy to realize a low-resistivity LLZ/LCO interface by the sintering process is to insert a coating layer such as LGVO at the interface. Another strategy will be to develop the interface at lower temperatures, which may require low-temperature process technology such as AD.

Conclusions

Click to copy section linkSection link copied!

We applied the concept of function sharing of a solid electrolyte to an oxide-based all-solid-state oxide battery (Ox-SSB) prepared by sintering processes similar to ML ceramic capacitors technology. We prepared an LLZ–LGVO ML solid electrolyte sheet where a 2 μm thick LGVO film was coated on one side of the LLZ substrate by AD. The ML obtained a total ionic conductivity of 6.9 × 10–4 S cm–1 at 25 °C and was stable up to 900 °C. These two solid electrolytes were well-densified through Ge diffusion and did not provide large interfacial resistance. MLs annealed at 900 °C (ML-900) were applied to develop Ox-SSBs (Li/ML-900/LCO) through sintering at 700 °C. The LGVO coating suppressed the increases in the resistance from the solid electrolyte and interfacial resistance induced by annealing by ca. 1/40. Additionally, the LGVO/LCO interfacial resistivity was decreased to as low as 145 Ω cm2 at 25 °C. Here, a 2 μm coating of LGVO is a tunable parameter in the field of ML ceramic capacitors. (43) Because the LLZ sheet has been successfully prepared by the tape casting method, (44) we can expect that the LLZ–LGVO ML will possibly be prepared through ML ceramic capacitors technology. Of course, the application of LGVO as a coating material may be another choice to develop better performance bulk-type Ox-SSBs (27) by reducing the Ohmic resistance in the composite electrode.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsami.2c10853.

  • EDX analysis of an LLZ–LGVO multilayer, reactivity between LGVO and LCO analyzed by Raman spectroscopy, cross-sectional SEM and EDX images of an LLZ/LGVO/LCO multilayer, DRT analysis of an EIS spectrum of an Ox-SSB (Li/ML-900/LCO), and EIS spectra of an Ox-SSB (Li/ML-900/LCO) (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Authors
    • Miyuki Sakakura - Department of Materials Design Innovation Engineering, Graduate School of Engineering, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8603, Japan
    • Kazutaka Mitsuishi - Research Center for Advanced Measurement and Characterization, National Institute for Materials Science, 1-2-1 Sengen, Tsukuba 305-0047, JapanOrcidhttps://orcid.org/0000-0002-9361-4057
    • Toyoki Okumura - Research Institute of Electrochemical Energy, National Institute of Advanced Industrial Science and Technology (AIST), 1-8-31 Midorigaoka, Ikeda, Osaka 563-8577, JapanOrcidhttps://orcid.org/0000-0003-0752-0941
    • Norikazu Ishigaki - Department of Materials Design Innovation Engineering, Graduate School of Engineering, Nagoya University, Furo-cho, Chikusa-ku, Nagoya 464-8603, Japan
  • Author Contributions

    The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

  • Funding

    This work was supported by ALCA-SPRING (JPMJAL1301) and in part by JSPS KAKENHI number JP19H05813 (Grant-in-Aid for Scientific Research on Innovation Areas “Interface IONICS”).

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

We thank Keisuke Shinoda and Tomoyuki Horie for TEM sample preparation at the National Institute for Materials Science (NIMS) Battery Research Platform. A part of this work was supported by the NIMS Electron Microscopy Analysis Station, Nanostructural Characterization Group.

References

Click to copy section linkSection link copied!

This article references 44 other publications.

  1. 1
    Bates, J. B.; Dudney, N. J.; Gruzalski, G. R.; Zuhr, R. A.; Choudhury, A.; Luck, C. F.; Robertson, J. D. Electrical Properties of Amorphous Lithium Electrolyte Thin Films. Solid State Ionics 1992, 53-56, 647654,  DOI: 10.1016/0167-2738(92)90442-r
  2. 2
    Bates, J. B.; Dudney, N. J.; Neudecker, B.; Ueda, A.; Evans, C. D. Thin-Film Lithium and Lithium-Ion Batteries. Solid State Ionics 2000, 135, 3345,  DOI: 10.1016/s0167-2738(00)00327-1
  3. 3
    https://www.tdk-electronics.tdk.com/en/ceracharge (accessed July 21, 2022).
  4. 4
    https://www.murata.com/en-eu/news/batteries/solid_state/2019/0626 (accessed July 21, 2022).
  5. 5
    Kim, K. H.; Iriyama, Y.; Yamamoto, K.; Kumazaki, S.; Asaka, T.; Tanabe, K.; Fisher, C. A.; Hirayama, T.; Murugan, R.; Ogumi, Z. Characterization of the interface between LiCoO2 and Li7La3Zr2O12 in an all-solid-state rechargeable lithium battery. J. Power Sources 2011, 196, 764767,  DOI: 10.1016/j.jpowsour.2010.07.073
  6. 6
    Iriyama, Y.; Wadaguchi, M.; Yoshida, K.; Yamamoto, Y.; Motoyama, M.; Yamamoto, T. 5V-Class Bulk-type All-Solid-State Rechargeable Lithium Batteries with Electrode-Solid Electrolyte Composite Electrodes Prepared by Aerosol Deposition. J. Power Sources 2018, 385, 5561,  DOI: 10.1016/j.jpowsour.2018.03.017
  7. 7
    Kato, T.; Yoshida, R.; Yamamoto, K.; Hirayama, T.; Motoyama, M.; West, W. C.; Iriyama, Y. Effects of Sintering Temperature on Interfacial Structure and Interfacial Resistance for All-Solid-State Rechargeable Lithium Batteries. J. Power Sources 2016, 325, 584590,  DOI: 10.1016/j.jpowsour.2016.06.068
  8. 8
    Murugan, R.; Thangadurai, V.; Weppner, W. Fast Lithium Ion Conduction in Garnet-Type Li7La3Zr2O12. Angew. Chem., Int. Ed. 2007, 46, 77787781,  DOI: 10.1002/anie.200701144
  9. 9
    Zhu, Y.; He, X.; Mo, Y. First Principles Study on Electrochemical and Chemical Stability of Solid Electrolyte-Electrode Interfaces in All-Solid-State Li-Ion Batteries. J. Mater. Chem. A 2016, 4, 32533266,  DOI: 10.1039/c5ta08574h
  10. 10
    Xu, L.; Li, J.; Deng, W.; Shuai, H.; Li, S.; Xu, Z.; Li, J.; Hou, H.; Peng, H.; Zou, G.; Ji, X. Garnet Solid Electrolyte for Advanced All-Solid-State Li Batteries. Adv. Energy Mater. 2021, 11, 2000648,  DOI: 10.1002/aenm.202000648
  11. 11
    Indu, M. S.; Alexander, G. V.; Sreejith, O. V.; Abraham, S. E.; Murugan, R. Lithium Garnet-Cathode Interfacial Chemistry: Inclusive Insights and Outlook Toward Practical Solid-State Lithium Metal Batteries. Mater. Today Energy 2021, 21, 100804,  DOI: 10.1016/j.mtener.2021.100804
  12. 12
    Sakuda, A.; Hayashi, A.; Tatsumisago, M. Interfacial Observation between LiCoO2 Electrode and Li2S–P2S5 Solid Electrolytes of All-Solid-State Lithium Secondary Batteries Using Transmission Electron Microscopy. Chem. Mater. 2010, 22, 949956,  DOI: 10.1021/cm901819c
  13. 13
    Zhang, W.; Richter, F. H.; Culver, S. P.; Leichtweiss, T.; Lozano, J. G.; Dietrich, C.; Bruce, P. G.; Zeier, W. G.; Janek, J. Degradation Mechanisms at the Li10GeP2S12/LiCoO2 Cathode Interface in an All-Solid-State Lithium-Ion Battery. ACS Appl. Mater. Interfaces 2018, 10, 2222622236,  DOI: 10.1021/acsami.8b05132
  14. 14
    Ohta, N.; Takada, K.; Zhang, L.; Ma, R.; Osada, M.; Sasaki, T. Enhancement of the High-Rate Capability of Solid-State Lithium Batteries by Nanoscale Interfacial Modification. Adv. Mater. 2006, 18, 22262229,  DOI: 10.1002/adma.200502604
  15. 15
    Haruyama, J.; Sodeyama, K.; Han, L.; Takada, K.; Tateyama, Y. Space-Charge Layer Effect at Interface between Oxide Cathode and Sulfide Electrolyte in All-Solid-State Lithium-Ion Battery. Chem. Mater. 2014, 26, 42484255,  DOI: 10.1021/cm5016959
  16. 16
    Zhu, Y.; He, X.; Mo, Y. Origin of Outstanding Stability in the Lithium Solid Electrolyte Materials: Insights from Thermodynamic Analyses Based on First-Principles Calculations. ACS Appl. Mater. Interfaces 2015, 7, 2368523693,  DOI: 10.1021/acsami.5b07517
  17. 17
    Fu, J. Fast Li+ Ion Conduction in Li2O-Al2O3-TiO2-SiO2-P2O2 Glass-Ceramics. J. Am. Ceram. Soc. 1997, 80, 19011903,  DOI: 10.1111/j.1151-2916.1997.tb03070.x
  18. 18
    Amiki, Y.; Sagane, F.; Yamamoto, K.; Hirayama, T.; Sudoh, M.; Motoyama, M.; Iriyama, Y. Electrochemical Properties of an All-Solid-State Lithium-Ion Battery with an In-Situ Formed electrode Material Grown from a Lithium Conductive Glass Ceramics Sheet. J. Power Sources 2013, 241, 583588,  DOI: 10.1016/j.jpowsour.2013.05.006
  19. 19
    Cheng, D.; Wynn, T. A.; Wang, X.; Wang, S.; Zhang, M.; Shimizu, R.; Bai, S.; Nguyen, H.; Fang, C.; Kim, M.-C.; Li, W.; Lu, B.; Kim, S.-J.; Meng, Y. S. Unveiling the Stable Nature of the Solid Electrolyte Interphase between Lithium Metal and LiPON via Cryogenic Electron Microscopy. Joule 2020, 4, 24842500,  DOI: 10.1016/j.joule.2020.08.013
  20. 20
    West, W. C.; Whitacre, J. F.; Lim, J. R. Chemical stability enhancement of lithium conducting solid electrolyte plates using sputtered LiPON thin films. J. Power Sources 2004, 126, 134138,  DOI: 10.1016/j.jpowsour.2003.08.030
  21. 21
    Sagane, F.; Ikeda, K.; Okita, K.; Sano, H.; Sakaebe, H.; Iriyama, Y. Effects of Current Densities on the Lithium Plating Morphology at a Lithium Phosphorus Oxynitride Glass Electrolyte/Copper Thin Film Interface. J. Power Sources 2013, 233, 3442,  DOI: 10.1016/j.jpowsour.2013.01.051
  22. 22
    Ohta, S.; Komagata, S.; Seki, J.; Saeki, T.; Morishita, S.; Asaoka, T. All-solid-state lithium ion battery using garnet-type oxide and Li3BO3 solid electrolytes fabricated by screen-printing. J. Power Sources 2013, 238, 5356,  DOI: 10.1016/j.jpowsour.2013.02.073
  23. 23
    Park, K.; Yu, B. C.; Jung, J. W.; Li, Y.; Zhou, W.; Gao, H.; Son, S.; Goodenough, J. B. Electrochemical Nature of the Cathode Interface for a Solid-State Lithium-Ion Battery: Interface between LiCoO2 and Garnet-Li7La3Zr2O12. Chem. Mater. 2016, 28, 80518059,  DOI: 10.1021/acs.chemmater.6b03870
  24. 24
    Shannon, R. D.; Taylor, B. E.; English, A. D.; Berzins, T. New Li Solid Electrolytes. Electrochim. Acta 1977, 22, 783796,  DOI: 10.1016/0013-4686(77)80035-2
  25. 25
    Okumura, T.; Takeuchi, T.; Kobayashi, H. All-solid-state lithium-ion battery using Li2.2C0.8B0.2O3 electrolyte. Solid State Ionics 2016, 288, 248252,  DOI: 10.1016/j.ssi.2016.01.045
  26. 26
    Han, F.; Yue, J.; Chen, C.; Zhao, N.; Fan, X.; Ma, Z.; Gao, T.; Wang, F.; Guo, X.; Wang, C. Interphase Engineering enabled All-Ceramic Lithium Battery. Joule 2018, 2, 497508,  DOI: 10.1016/j.joule.2018.02.007
  27. 27
    Okumura, T.; Takeuchi, T.; Kobayashi, H. All-Solid-State Batteries with LiCoO2-Type Electrodes: Realization of an Impurity-Free Interface by Utilizing a Cosinterable Li3.5Ge0.5V0.5O4 Electrolyte. ACS Appl. Energy Mater. 2021, 4, 3034,  DOI: 10.1021/acsaem.0c02785
  28. 28
    Akedo, J.; Lebedev, M. Microstructure and Electrical Properties of Lead Zirconate Titanate (Pb(Zr52/Ti48)O3) Thick Films Deposited by Aerosol Deposition Method. Jpn. J. Appl. Phys. 1999, 38, 53975401,  DOI: 10.1143/jjap.38.5397
  29. 29
    Fuchita, E.; Tokizaki, E.; Ozawa, E.; Sakka, Y. Formation of Zirconia Films by the Aerosol Gas Deposition Method (By Jetting of Positive Charged Powder). J. Jpn. Soc. Powder Powder Metall. 2011, 58, 463472,  DOI: 10.2497/jjspm.58.463
  30. 30
    Iwasaki, S.; Hamanaka, T.; Yamakawa, T.; West, W. C.; Yamamoto, K.; Motoyama, M.; Hirayama, T.; Iriyama, Y. Preparation of thick-film LiNi1/3Co1/3Mn1/3O2 electrodes by aerosol deposition and its application to all-solid-state batteries. J. Power Sources 2014, 272, 10861090,  DOI: 10.1016/j.jpowsour.2014.09.038
  31. 31
    Kato, T.; Iwasaki, S.; Ishii, Y.; Motoyama, M.; West, W. C.; Yamamoto, Y.; Iriyama, Y. Preparation of thick-film electrode-solid electrolyte composites on Li7La3Zr2O12 and their electrochemical properties. J. Power Sources 2016, 303, 6572,  DOI: 10.1016/j.jpowsour.2015.10.101
  32. 32
    Motoyama, M.; Tanaka, Y.; Yamamoto, T.; Tsuchimine, N.; Kobayashi, S.; Iriyama, Y. The Active Interface of Ta-Doped Li7La3Zr2O12 for Li Plating/Stripping Revealed by Acid Aqueous Etching. ACS Appl. Energy Mater. 2019, 2, 67206731,  DOI: 10.1021/acsaem.9b01193
  33. 33
    Motoyama, M.; Iwasaki, H.; Sakakura, M.; Yamamoto, T.; Iriyama, Y. Synthesis of LiCoO2 particles with tunable sizes by a urea-based-homogeneous-precipitation method. Int. J. Mater. Res. 2020, 111, 347355,  DOI: 10.1515/ijmr-2020-1110411
  34. 34
    Abrahams, I.; Bruce, P. G. Defect Clustering in the Superionic Conductor Lithium Germanium Vanadate. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 1991, 47, 696701,  DOI: 10.1107/s0108768191004548
  35. 35
    Kuwano, J.; West, A. R. New Li+ ion conductors in the system, Li4GeO4-Li3VO4. Mater. Res. Bull. 1980, 15, 16611667,  DOI: 10.1016/0025-5408(80)90249-4
  36. 36
    Huang, M.; Xu, W.; Shen, Y.; Lin, Y.-H.; Nan, C.-W. X-ray absorption near-edge spectroscopy study on Ge-doped Li7La3Zr2O12: enhanced ionic conductivity and defect chemistry. Electrochim. Acta 2014, 115, 581586,  DOI: 10.1016/j.electacta.2013.11.020
  37. 37
    Watanabe, M.; Williams, D. B. The quantitative analysis of thin specimens: a review of progress from the Cliff-Lorimer to the new zeta-factor methods. J. Microsc. 2006, 221, 89109,  DOI: 10.1111/j.1365-2818.2006.01549.x
  38. 38
    Rangasamy, E.; Wolfenstine, J.; Sakamoto, J. The role of Al and Li concentration on the formation of cubic garnet solid electrolyte of nominal composition Li7La3Zr2O12. Solid State Ionics 2012, 206, 2832,  DOI: 10.1016/j.ssi.2011.10.022
  39. 39
    Inaba, M.; Iriyama, Y.; Ogumi, Z.; Todzuka, Y.; Tasaka, A. Raman study of layered rock-salt LiCoO2 and its electrochemical lithium deintercalation. J. Raman Spectrosc. 1997, 28, 613617,  DOI: 10.1002/(sici)1097-4555(199708)28:8<613::aid-jrs138>3.0.co;2-t
  40. 40
    Reimers, J. N.; Dahn, J. R. Electrochemical and In Situ X-Ray Diffraction Studies of Lithium Intercalation in Li x CoO2. J. Electrochem. Soc. 1992, 139, 20912097,  DOI: 10.1149/1.2221184
  41. 41
    Kato, T.; Hamanaka, T.; Yamamoto, K.; Hirayama, T.; Sagane, F.; Motoyama, M.; Iriyama, Y. In-situ Li7La3Zr2O12/LiCoO2 interface modification for advanced all-solid-state battery. J. Power Sources 2014, 260, 292298,  DOI: 10.1016/j.jpowsour.2014.02.102
  42. 42
    Richards, W. D.; Miara, L. J.; Wang, Y.; Kim, J. C.; Ceder, G. Interface Stability in Solid-State Batteries. Chem. Mater. 2016, 28, 266273,  DOI: 10.1021/acs.chemmater.5b04082
  43. 43
    Kishi, H.; Mizuno, Y.; Chazono, H. Base-Metal Electrode-Multilayer Ceramic Capacitors: Past, Present and Future Perspectives. Jpn. J. Appl. Phys. 2003, 42, 115,  DOI: 10.1143/jjap.42.1
  44. 44
    Hitz, G. T.; McOwen, D. W.; Zhang, L.; Ma, Z.; Fu, Z.; Wen, Y.; Gong, Y.; Dai, J.; Hamann, T. R.; Hu, L.; Wachsman, E. D. High-Rate Lithium Cycling in a Scalable Trilayer Li-Garnet-Electrolyte Architecture. Mater. Today 2019, 22, 5057,  DOI: 10.1016/j.mattod.2018.04.004

Cited By

Click to copy section linkSection link copied!

This article is cited by 6 publications.

  1. Myeong Jun Joo, Minseong Kim, Sujong Chae, Minseong Ko, Yong Joon Park. Additive-Derived Surface Modification of Cathodes in All-Solid-State Batteries: The Effect of Lithium Difluorophosphate- and Lithium Difluoro(oxalato)borate-Derived Coating Layers. ACS Applied Materials & Interfaces 2023, 15 (51) , 59389-59402. https://doi.org/10.1021/acsami.3c12858
  2. Yuzhou Bai, Wenqin Ma, Wujie Dong, Yingkang Wu, Xue Wang, Fuqiang Huang. In-Situ-Polymerized 1,3-Dioxolane Solid-State Electrolyte with Space-Confined Plasticizers for High-Voltage and Robust Li/LiCoO2 Batteries. ACS Applied Materials & Interfaces 2023, 15 (22) , 26834-26842. https://doi.org/10.1021/acsami.3c04234
  3. Chunxiang Xian, Qiyue Wang, Yang Xia, Feng Cao, Shenghui Shen, Yongqi Zhang, Minghua Chen, Yu Zhong, Jun Zhang, Xinping He, Xinhui Xia, Wenkui Zhang, Jiangping Tu. Solid‐State Electrolytes in Lithium–Sulfur Batteries: Latest Progresses and Prospects. Small 2023, 19 (24) https://doi.org/10.1002/smll.202208164
  4. Deidre Wolff, Svenja Weber, Tobias Graumann, Stefan Zebrowski, Nils Mainusch, Nikolas Dilger, Felipe Cerdas, Sabrina Zellmer. An Environmental and Technical Evaluation of Vacuum-Based Thin Film Technologies: Lithium Niobate Coated Cathode Active Material for Use in All-Solid-State Battery Cells. Energies 2023, 16 (3) , 1278. https://doi.org/10.3390/en16031278
  5. Miyuki Sakakura, Yasutoshi Iriyama. Development of oxide-based all-solid-state batteries using aerosol deposition. Journal of Asian Ceramic Societies 2023, 11 (1) , 1-10. https://doi.org/10.1080/21870764.2022.2163080
  6. Yuan Yang, Nai-Fang Hu, Yong-Cheng Jin, Jun Ma, Guang-Lei Cui, , . Research advance of lithium-rich cathode materials in all-solid-state lithium batteries. Acta Physica Sinica 2023, 72 (11) , 118801. https://doi.org/10.7498/aps.72.20230258

ACS Applied Materials & Interfaces

Cite this: ACS Appl. Mater. Interfaces 2022, 14, 43, 48547–48557
Click to copy citationCitation copied!
https://doi.org/10.1021/acsami.2c10853
Published October 3, 2022

Copyright © 2022 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

2447

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. (a) FE-SEM image of LGVO particles for AD. The inset shows a magnified SEM image of a particle. (b) XRD patterns of the LGVO particles (black) and reported reference data (red). (c) Nyquist plots of a Au/LGVO/Au symmetrical cell at 25 °C. The inset displays a magnified image of a higher frequency region. The numbers in these plots indicate data points measured at 10n Hz.

    Figure 2

    Figure 2. (a–d) Surface and (e–f) cross-sectional FE-SEM images of (a,e) ML-AP, (b,f) ML-700, (c,g) ML-900, and (d,h) ML-1050. The insets in (a–d) display optical images of the MLs.

    Figure 3

    Figure 3. Cross-sectional TEM images of (a) ML-AP and (b) ML-900. EDX measurements around the LLZ/LGVO interface of (c) ML-AP and (d) ML-900. The EDX signals were from O Kα, V Kα, Ge Kα, Zr Lα, La Lα, and Ta Lα. The tables in Figure 2c,d summarize the mass %, σ, and at. % evaluated in the LLZ region in the red frame.

    Figure 4

    Figure 4. XRD patterns of ML-AP (blue), ML-700 (orange), ML-900 (red), and ML-1050 (purple). Additionally, XRD patterns of the LLZ substrate after polishing with SiC sandpaper (black) and LGVO powder (green) are shown as references. Diffraction peaks marked by * are assigned to La2Zr2O7.

    Figure 5

    Figure 5. (a) Nyquist plots of ML-AP (blue), ML-700 (orange), ML-900 (red), ML-1050 (purple), and the LLZ substrate (green) measured at 25 °C. Magnified Nyquist plots in a higher frequency region of the (b) LLZ substrate, (c) ML-AP, (d) ML-700, (e) ML-900, and (f) ML-1050. The numbers in each plot indicate data points measured at 10n Hz. The thickness of each substrate is used to normalize all the resistances.

    Figure 6

    Figure 6. Variation in the discharge capacity and total interfacial resistance (Rint) of an Ox-SSB (Li/ML-900/LCO) with the charge–discharge reactions measured at 60 and 25 °C. The charge current density was fixed at 2.9 μA cm–2 in the 1st–36th cycles and 5 μA cm–2 in the 37th–137th cycles, and the discharge current density was varied from 2.9 to 286 μA cm–2 during the 1st–36th cycles and 5–500 μA cm–2 during the 37th–91st cycles and fixed at 50 μA cm–2 during the 92nd–137th cycles. The Rint values were measured at 4.0 V after charge–discharge reactions in the 5th, 25th, 36th, 86th, and 136th cycles.

    Figure 7

    Figure 7. Charge–discharge curves of an Ox-SSB (Li/ML-900/LCO) with 700 °C annealing in the (a) 1st, 6th, 11th, 16th, and 21st cycles and (b) 26th, 31st, and 36th cycles, where the charge current density was fixed at 2.9 μA cm–2 and discharge current density was 2.9 μA cm–2 (1st and 26th), 29 μA cm–2 (6th and 31st), 57 μA cm–2 (11th and 36th), 143 μA cm–2 (16th), and 286 μA cm–2 (21st). Nyquist plots at 3.0 and 4.0 V after the (c) 5th cycle and (d) 25th cycle, where Rint denotes the sum of the interfacial resistances at Li/LLZ and LGVO/LCO. Data were measured at 60 °C in (a,c) and at 25 °C in (b,d). The numbers in (c,d) indicate data points measured at 10n Hz.

    Figure 8

    Figure 8. (a) Charge–discharge curves at 60 °C of an Ox-SSB (Li/LLZ/LCO) with 700 °C annealing in the 1st–5th cycles under a fixed charge–discharge current density of 5 μA cm–2. (b) Nyquist plots at 60 °C of a nonannealed Ox-SSB (Li/LLZ) measured at 3.0 V (blue) and 4.0 V (red) after the 5th cycle. (c) Charge–discharge curves at 25 °C of a nonannealed Ox-SSB (Li/LLZ/LCO) in the 6th–10th cycles under a fixed charge–discharge current density of 5 μA cm–2. (d) Nyquist plots at 25 °C of an Ox-SSB (Li/LLZ/LCO) without annealing measured at 3.0 V (blue) and 4.0 V (red) after the 10th cycle. (e) Charge–discharge curves at 60 °C of an Ox-SSB (Li/LLZ/LCO) without annealing in the 1st–5th cycles under a fixed charge–discharge current density of 5 μA cm–2. (f) Nyquist plots at 60 °C of an Ox-SSB (Li/LLZ) without annealing measured at 3.0 V (blue) and 4.0 V (red) after the 5th cycle. (g) Schematic image of the interface properties with LGVO coating and annealing effects.

  • References


    This article references 44 other publications.

    1. 1
      Bates, J. B.; Dudney, N. J.; Gruzalski, G. R.; Zuhr, R. A.; Choudhury, A.; Luck, C. F.; Robertson, J. D. Electrical Properties of Amorphous Lithium Electrolyte Thin Films. Solid State Ionics 1992, 53-56, 647654,  DOI: 10.1016/0167-2738(92)90442-r
    2. 2
      Bates, J. B.; Dudney, N. J.; Neudecker, B.; Ueda, A.; Evans, C. D. Thin-Film Lithium and Lithium-Ion Batteries. Solid State Ionics 2000, 135, 3345,  DOI: 10.1016/s0167-2738(00)00327-1
    3. 3
      https://www.tdk-electronics.tdk.com/en/ceracharge (accessed July 21, 2022).
    4. 4
      https://www.murata.com/en-eu/news/batteries/solid_state/2019/0626 (accessed July 21, 2022).
    5. 5
      Kim, K. H.; Iriyama, Y.; Yamamoto, K.; Kumazaki, S.; Asaka, T.; Tanabe, K.; Fisher, C. A.; Hirayama, T.; Murugan, R.; Ogumi, Z. Characterization of the interface between LiCoO2 and Li7La3Zr2O12 in an all-solid-state rechargeable lithium battery. J. Power Sources 2011, 196, 764767,  DOI: 10.1016/j.jpowsour.2010.07.073
    6. 6
      Iriyama, Y.; Wadaguchi, M.; Yoshida, K.; Yamamoto, Y.; Motoyama, M.; Yamamoto, T. 5V-Class Bulk-type All-Solid-State Rechargeable Lithium Batteries with Electrode-Solid Electrolyte Composite Electrodes Prepared by Aerosol Deposition. J. Power Sources 2018, 385, 5561,  DOI: 10.1016/j.jpowsour.2018.03.017
    7. 7
      Kato, T.; Yoshida, R.; Yamamoto, K.; Hirayama, T.; Motoyama, M.; West, W. C.; Iriyama, Y. Effects of Sintering Temperature on Interfacial Structure and Interfacial Resistance for All-Solid-State Rechargeable Lithium Batteries. J. Power Sources 2016, 325, 584590,  DOI: 10.1016/j.jpowsour.2016.06.068
    8. 8
      Murugan, R.; Thangadurai, V.; Weppner, W. Fast Lithium Ion Conduction in Garnet-Type Li7La3Zr2O12. Angew. Chem., Int. Ed. 2007, 46, 77787781,  DOI: 10.1002/anie.200701144
    9. 9
      Zhu, Y.; He, X.; Mo, Y. First Principles Study on Electrochemical and Chemical Stability of Solid Electrolyte-Electrode Interfaces in All-Solid-State Li-Ion Batteries. J. Mater. Chem. A 2016, 4, 32533266,  DOI: 10.1039/c5ta08574h
    10. 10
      Xu, L.; Li, J.; Deng, W.; Shuai, H.; Li, S.; Xu, Z.; Li, J.; Hou, H.; Peng, H.; Zou, G.; Ji, X. Garnet Solid Electrolyte for Advanced All-Solid-State Li Batteries. Adv. Energy Mater. 2021, 11, 2000648,  DOI: 10.1002/aenm.202000648
    11. 11
      Indu, M. S.; Alexander, G. V.; Sreejith, O. V.; Abraham, S. E.; Murugan, R. Lithium Garnet-Cathode Interfacial Chemistry: Inclusive Insights and Outlook Toward Practical Solid-State Lithium Metal Batteries. Mater. Today Energy 2021, 21, 100804,  DOI: 10.1016/j.mtener.2021.100804
    12. 12
      Sakuda, A.; Hayashi, A.; Tatsumisago, M. Interfacial Observation between LiCoO2 Electrode and Li2S–P2S5 Solid Electrolytes of All-Solid-State Lithium Secondary Batteries Using Transmission Electron Microscopy. Chem. Mater. 2010, 22, 949956,  DOI: 10.1021/cm901819c
    13. 13
      Zhang, W.; Richter, F. H.; Culver, S. P.; Leichtweiss, T.; Lozano, J. G.; Dietrich, C.; Bruce, P. G.; Zeier, W. G.; Janek, J. Degradation Mechanisms at the Li10GeP2S12/LiCoO2 Cathode Interface in an All-Solid-State Lithium-Ion Battery. ACS Appl. Mater. Interfaces 2018, 10, 2222622236,  DOI: 10.1021/acsami.8b05132
    14. 14
      Ohta, N.; Takada, K.; Zhang, L.; Ma, R.; Osada, M.; Sasaki, T. Enhancement of the High-Rate Capability of Solid-State Lithium Batteries by Nanoscale Interfacial Modification. Adv. Mater. 2006, 18, 22262229,  DOI: 10.1002/adma.200502604
    15. 15
      Haruyama, J.; Sodeyama, K.; Han, L.; Takada, K.; Tateyama, Y. Space-Charge Layer Effect at Interface between Oxide Cathode and Sulfide Electrolyte in All-Solid-State Lithium-Ion Battery. Chem. Mater. 2014, 26, 42484255,  DOI: 10.1021/cm5016959
    16. 16
      Zhu, Y.; He, X.; Mo, Y. Origin of Outstanding Stability in the Lithium Solid Electrolyte Materials: Insights from Thermodynamic Analyses Based on First-Principles Calculations. ACS Appl. Mater. Interfaces 2015, 7, 2368523693,  DOI: 10.1021/acsami.5b07517
    17. 17
      Fu, J. Fast Li+ Ion Conduction in Li2O-Al2O3-TiO2-SiO2-P2O2 Glass-Ceramics. J. Am. Ceram. Soc. 1997, 80, 19011903,  DOI: 10.1111/j.1151-2916.1997.tb03070.x
    18. 18
      Amiki, Y.; Sagane, F.; Yamamoto, K.; Hirayama, T.; Sudoh, M.; Motoyama, M.; Iriyama, Y. Electrochemical Properties of an All-Solid-State Lithium-Ion Battery with an In-Situ Formed electrode Material Grown from a Lithium Conductive Glass Ceramics Sheet. J. Power Sources 2013, 241, 583588,  DOI: 10.1016/j.jpowsour.2013.05.006
    19. 19
      Cheng, D.; Wynn, T. A.; Wang, X.; Wang, S.; Zhang, M.; Shimizu, R.; Bai, S.; Nguyen, H.; Fang, C.; Kim, M.-C.; Li, W.; Lu, B.; Kim, S.-J.; Meng, Y. S. Unveiling the Stable Nature of the Solid Electrolyte Interphase between Lithium Metal and LiPON via Cryogenic Electron Microscopy. Joule 2020, 4, 24842500,  DOI: 10.1016/j.joule.2020.08.013
    20. 20
      West, W. C.; Whitacre, J. F.; Lim, J. R. Chemical stability enhancement of lithium conducting solid electrolyte plates using sputtered LiPON thin films. J. Power Sources 2004, 126, 134138,  DOI: 10.1016/j.jpowsour.2003.08.030
    21. 21
      Sagane, F.; Ikeda, K.; Okita, K.; Sano, H.; Sakaebe, H.; Iriyama, Y. Effects of Current Densities on the Lithium Plating Morphology at a Lithium Phosphorus Oxynitride Glass Electrolyte/Copper Thin Film Interface. J. Power Sources 2013, 233, 3442,  DOI: 10.1016/j.jpowsour.2013.01.051
    22. 22
      Ohta, S.; Komagata, S.; Seki, J.; Saeki, T.; Morishita, S.; Asaoka, T. All-solid-state lithium ion battery using garnet-type oxide and Li3BO3 solid electrolytes fabricated by screen-printing. J. Power Sources 2013, 238, 5356,  DOI: 10.1016/j.jpowsour.2013.02.073
    23. 23
      Park, K.; Yu, B. C.; Jung, J. W.; Li, Y.; Zhou, W.; Gao, H.; Son, S.; Goodenough, J. B. Electrochemical Nature of the Cathode Interface for a Solid-State Lithium-Ion Battery: Interface between LiCoO2 and Garnet-Li7La3Zr2O12. Chem. Mater. 2016, 28, 80518059,  DOI: 10.1021/acs.chemmater.6b03870
    24. 24
      Shannon, R. D.; Taylor, B. E.; English, A. D.; Berzins, T. New Li Solid Electrolytes. Electrochim. Acta 1977, 22, 783796,  DOI: 10.1016/0013-4686(77)80035-2
    25. 25
      Okumura, T.; Takeuchi, T.; Kobayashi, H. All-solid-state lithium-ion battery using Li2.2C0.8B0.2O3 electrolyte. Solid State Ionics 2016, 288, 248252,  DOI: 10.1016/j.ssi.2016.01.045
    26. 26
      Han, F.; Yue, J.; Chen, C.; Zhao, N.; Fan, X.; Ma, Z.; Gao, T.; Wang, F.; Guo, X.; Wang, C. Interphase Engineering enabled All-Ceramic Lithium Battery. Joule 2018, 2, 497508,  DOI: 10.1016/j.joule.2018.02.007
    27. 27
      Okumura, T.; Takeuchi, T.; Kobayashi, H. All-Solid-State Batteries with LiCoO2-Type Electrodes: Realization of an Impurity-Free Interface by Utilizing a Cosinterable Li3.5Ge0.5V0.5O4 Electrolyte. ACS Appl. Energy Mater. 2021, 4, 3034,  DOI: 10.1021/acsaem.0c02785
    28. 28
      Akedo, J.; Lebedev, M. Microstructure and Electrical Properties of Lead Zirconate Titanate (Pb(Zr52/Ti48)O3) Thick Films Deposited by Aerosol Deposition Method. Jpn. J. Appl. Phys. 1999, 38, 53975401,  DOI: 10.1143/jjap.38.5397
    29. 29
      Fuchita, E.; Tokizaki, E.; Ozawa, E.; Sakka, Y. Formation of Zirconia Films by the Aerosol Gas Deposition Method (By Jetting of Positive Charged Powder). J. Jpn. Soc. Powder Powder Metall. 2011, 58, 463472,  DOI: 10.2497/jjspm.58.463
    30. 30
      Iwasaki, S.; Hamanaka, T.; Yamakawa, T.; West, W. C.; Yamamoto, K.; Motoyama, M.; Hirayama, T.; Iriyama, Y. Preparation of thick-film LiNi1/3Co1/3Mn1/3O2 electrodes by aerosol deposition and its application to all-solid-state batteries. J. Power Sources 2014, 272, 10861090,  DOI: 10.1016/j.jpowsour.2014.09.038
    31. 31
      Kato, T.; Iwasaki, S.; Ishii, Y.; Motoyama, M.; West, W. C.; Yamamoto, Y.; Iriyama, Y. Preparation of thick-film electrode-solid electrolyte composites on Li7La3Zr2O12 and their electrochemical properties. J. Power Sources 2016, 303, 6572,  DOI: 10.1016/j.jpowsour.2015.10.101
    32. 32
      Motoyama, M.; Tanaka, Y.; Yamamoto, T.; Tsuchimine, N.; Kobayashi, S.; Iriyama, Y. The Active Interface of Ta-Doped Li7La3Zr2O12 for Li Plating/Stripping Revealed by Acid Aqueous Etching. ACS Appl. Energy Mater. 2019, 2, 67206731,  DOI: 10.1021/acsaem.9b01193
    33. 33
      Motoyama, M.; Iwasaki, H.; Sakakura, M.; Yamamoto, T.; Iriyama, Y. Synthesis of LiCoO2 particles with tunable sizes by a urea-based-homogeneous-precipitation method. Int. J. Mater. Res. 2020, 111, 347355,  DOI: 10.1515/ijmr-2020-1110411
    34. 34
      Abrahams, I.; Bruce, P. G. Defect Clustering in the Superionic Conductor Lithium Germanium Vanadate. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 1991, 47, 696701,  DOI: 10.1107/s0108768191004548
    35. 35
      Kuwano, J.; West, A. R. New Li+ ion conductors in the system, Li4GeO4-Li3VO4. Mater. Res. Bull. 1980, 15, 16611667,  DOI: 10.1016/0025-5408(80)90249-4
    36. 36
      Huang, M.; Xu, W.; Shen, Y.; Lin, Y.-H.; Nan, C.-W. X-ray absorption near-edge spectroscopy study on Ge-doped Li7La3Zr2O12: enhanced ionic conductivity and defect chemistry. Electrochim. Acta 2014, 115, 581586,  DOI: 10.1016/j.electacta.2013.11.020
    37. 37
      Watanabe, M.; Williams, D. B. The quantitative analysis of thin specimens: a review of progress from the Cliff-Lorimer to the new zeta-factor methods. J. Microsc. 2006, 221, 89109,  DOI: 10.1111/j.1365-2818.2006.01549.x
    38. 38
      Rangasamy, E.; Wolfenstine, J.; Sakamoto, J. The role of Al and Li concentration on the formation of cubic garnet solid electrolyte of nominal composition Li7La3Zr2O12. Solid State Ionics 2012, 206, 2832,  DOI: 10.1016/j.ssi.2011.10.022
    39. 39
      Inaba, M.; Iriyama, Y.; Ogumi, Z.; Todzuka, Y.; Tasaka, A. Raman study of layered rock-salt LiCoO2 and its electrochemical lithium deintercalation. J. Raman Spectrosc. 1997, 28, 613617,  DOI: 10.1002/(sici)1097-4555(199708)28:8<613::aid-jrs138>3.0.co;2-t
    40. 40
      Reimers, J. N.; Dahn, J. R. Electrochemical and In Situ X-Ray Diffraction Studies of Lithium Intercalation in Li x CoO2. J. Electrochem. Soc. 1992, 139, 20912097,  DOI: 10.1149/1.2221184
    41. 41
      Kato, T.; Hamanaka, T.; Yamamoto, K.; Hirayama, T.; Sagane, F.; Motoyama, M.; Iriyama, Y. In-situ Li7La3Zr2O12/LiCoO2 interface modification for advanced all-solid-state battery. J. Power Sources 2014, 260, 292298,  DOI: 10.1016/j.jpowsour.2014.02.102
    42. 42
      Richards, W. D.; Miara, L. J.; Wang, Y.; Kim, J. C.; Ceder, G. Interface Stability in Solid-State Batteries. Chem. Mater. 2016, 28, 266273,  DOI: 10.1021/acs.chemmater.5b04082
    43. 43
      Kishi, H.; Mizuno, Y.; Chazono, H. Base-Metal Electrode-Multilayer Ceramic Capacitors: Past, Present and Future Perspectives. Jpn. J. Appl. Phys. 2003, 42, 115,  DOI: 10.1143/jjap.42.1
    44. 44
      Hitz, G. T.; McOwen, D. W.; Zhang, L.; Ma, Z.; Fu, Z.; Wen, Y.; Gong, Y.; Dai, J.; Hamann, T. R.; Hu, L.; Wachsman, E. D. High-Rate Lithium Cycling in a Scalable Trilayer Li-Garnet-Electrolyte Architecture. Mater. Today 2019, 22, 5057,  DOI: 10.1016/j.mattod.2018.04.004
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsami.2c10853.

    • EDX analysis of an LLZ–LGVO multilayer, reactivity between LGVO and LCO analyzed by Raman spectroscopy, cross-sectional SEM and EDX images of an LLZ/LGVO/LCO multilayer, DRT analysis of an EIS spectrum of an Ox-SSB (Li/ML-900/LCO), and EIS spectra of an Ox-SSB (Li/ML-900/LCO) (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.