Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect
ACS Publications. Most Trusted. Most Cited. Most Read
Drastic Reduction of the Solid Electrolyte–Electrode Interface Resistance via Annealing in Battery Form
My Activity

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

Drastic Reduction of the Solid Electrolyte–Electrode Interface Resistance via Annealing in Battery Form
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

ACS Applied Materials & Interfaces

Cite this: ACS Appl. Mater. Interfaces 2022, 14, 2, 2703–2710
Click to copy citationCitation copied!
https://doi.org/10.1021/acsami.1c17945
Published January 6, 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!

The origin of electrical resistance at the interface between the positive electrode and solid electrolyte of an all-solid-state Li battery has not been fully determined. It is well known that the interface resistance increases when the electrode surface is exposed to air. However, an effective method of reducing this resistance has not been developed. This report demonstrates that drastic reduction of the resistance is achievable by annealing the entire battery cell. Exposing the LiCoO2 positive electrode surface to H2O vapor increases the resistance by more than 10 times (to greater than 136 Ω cm2). The magnitude can be reduced to the initial value (10.3 Ω cm2) by annealing the sample in a battery form. First-principles calculations reveal that the protons incorporated into the LiCoO2 structure are spontaneously deintercalated during annealing to restore the low-resistance interface. These results provide fundamental insights into the fabrication of high-performance all-solid-state Li batteries.

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!

All-solid-state Li batteries have become promising next-generation batteries owing to their high energy density and charge–discharge cyclability. (1) However, the large resistance at the interface between the solid electrolyte and positive electrode (electrolyte–electrode interface resistance) considerably deteriorates their performance. (2,3) In the proposed mechanisms of the increased interface resistance, one of the most plausible origins of the increased interface resistance is the chemical side reaction at the interface. (4−10) The atmospheric gas adsorption on the electrode surface significantly deteriorates the interface resistance, (11,12) and the battery capacity and cyclic performance are degraded. (13,14) Although there have been many attempts to decrease the interface resistance, such as by introducing oxide buffer layers, (15,16) no method can restore the resistance to ∼10 Ω cm2, which is close to the lowest solid-electrolyte–electrode interface resistance reported. (17) Establishing a strategy for recovering the low interface resistance is critically important for the development of all-solid-state Li batteries.
Thin-film Li batteries provide suitable platforms for studying interfacial phenomena. They enable the quantitative investigation of interface resistances by defining the areas and crystal structures of electrolyte–electrode interfaces. (18,19) In vacuo fabrication techniques enable the fabrication of electrolyte–electrode interfaces with very low resistance values. (17) Such low-resistance interfaces will enable the investigation of factors, such as gas exposure, which could affect the interface phenomena by manipulating them individually. Such bottom-up studies will lead to an advanced foundation of solid-state physics that is combined with electrochemistry.
This paper reports the recovery of a low interface resistance and elucidates its reduction mechanism. First, we demonstrate that, among the different gas species present in air, only H2O vapor strongly degrades the Li3PO4–LiCoO2 interface and drastically increases its resistance. Next, we show that the low interface resistance can be recovered by annealing the sample in a battery form (after depositing the negative electrode). Finally, we discuss the low resistance recovery atomistic mechanism determined from the results of structural analysis and first-principles calculations. We emphasize that the protons in the LiCoO2 structure play an important role in the recovery process.

Results

Click to copy section linkSection link copied!

Fabrication of Gas-Exposed Batteries and Their Performance Characteristics

We fabricated the thin-film batteries consisting of an Li negative electrode with a thickness (t) of 1 μm, an amorphous Li3PO4 solid electrolyte (t = 1 μm), an LiCoO2 positive electrode (t = 40 nm), an Au current collector (t = 100 nm), and an Al2O3 (0001) substrate (Figure S1). The interfaces between the Li3PO4 and LiCoO2 layers were intentionally exposed to air, N2, O2, CO2, H2, and H2O gas (hereafter referred to as the air-, N2-, O2-, CO2-, H2-, and H2O-exposed batteries, respectively) for 30 min. We also prepared a nonexposed battery (in vacuo) as a reference (see the Methods section for the details of the fabrication procedure).
Figure 1a,b compares the CV curves of the in vacuo and air-exposed thin-film batteries. The in vacuo battery exhibits sharp current peaks at 3.94 and 3.88 V vs Li+/Li, corresponding to the extraction and insertion of Li ions, respectively. In contrast, the air-exposed battery generates relatively weak peaks, indicating that exposing the electrode surface to air strongly degrades the battery performance.

Figure 1

Figure 1. Electrochemical characteristics of the (a, c, and e) in vacuo and (b, d, and f) air-exposed batteries. cyclic voltammetry (CV) curves of the (a) in vacuo and (b) air-exposed batteries. The voltage was swept at a rate of 1 mV·s–1 from 3.0 to 4.3 V vs Li+/Li. The displayed curves correspond to the 20th charge–discharge cycle. Impedance spectra of the (c) in vacuo and (d) air-exposed batteries recorded at voltages of 3.8 and 4.2 V vs Li+/Li. The numbers n in the plots indicate frequencies (10n Hz). Low-frequency regions of the impedance spectra recorded for the (e) in vacuo and (f) air-exposed batteries. The fitting data obtained for the equivalent circuit at 4.2 V vs Li+/Li data (Table S1) are shown as colored semicircles. The light-green semicircle denotes the resistance of Li ion conduction in the Li3PO4 solid electrolyte. The gray semicircle represents the Li3PO4–LiCoO2 interface resistance (10.9 Ω·cm2 for the in vacuo battery and 200 Ω·cm2 for the air-exposed battery).

The AC impedance spectra obtained for both batteries contain clear semicircles in the frequency range of 102–105 Hz (Figure 1c,d). Because these semicircles are observed both in the charged and discharged states (4.2 and 3.8 V vs Li+/Li), they likely originate from the impedance of the Li3PO4 solid electrolyte. The ionic conductivity of Li3PO4 estimated from the impedance spectra is 4.9 × 10–7 S cm–1, which is in good agreement with the previously reported values. (20) In addition to the semicircle in the high-frequency region, a semicircle associated with the electrolyte–electrode interface is observed in the 101–102 Hz frequency range. The fit of the semicircle in Figure 1e,f results in an interface resistance of 10.9 Ω·cm2 for the in vacuo battery and 200 Ω cm2 for the air-exposed battery (see the detailed analysis in Figure S2). The increase in the interface resistance observed after air exposure is consistent with the results obtained by Iriyama et al. (11)
Next, we discuss the effects of exposing the electrode surfaces to pure gases (N2, O2, CO2, H2, and H2O) on the interface resistance. Initially, we expected that these gas molecules would react with Li ions in the LiCoO2 lattice to form a resistive layer on the electrode surface. However, a typical interface resistance of the N2-exposed batteries is 9.6 Ω·cm2, which is very close to that of the in vacuo battery. In addition, comparable resistance values (∼10 Ω·cm2) were obtained for the O2-, CO2-, and H2-exposed batteries, indicating that N2, O2, CO2, and H2 gases did not degrade the interfaces (Table 1; Figure S3d–l).
Table 1. Interface Resistances of the Gas-Exposed Batteries
 exposed gas species
 in vacuoairN2CO2O2H2H2O
interface resistance at 4.2 V vs Li+/Li (Ω·cm2)10.92009.68.316.513.4>136
In contrast, H2O vapor exposure drastically degrades the battery performance. The peak current in the corresponding CV curve is ∼20 nA (10 mA/cm2, Figure 2a) at a voltage of 3.9 V vs Li+/Li, which is much smaller than the magnitude obtained for the in vacuo battery (approximately 200 nA (100 mA/cm2); see Figure 1a). It was difficult to determine the interface resistance of the H2O-exposed battery accurately because of the strong electrode degradation. Therefore, we speculate that the interface resistance is underestimated. We estimated its value to be greater than 136 Ω·cm2, which was at least 12 times larger than that of the in vacuo interface (Figure S4). H2O is the only species that increases the interface resistance; therefore, it is very important to recover the deteriorated interface after H2O exposure. (21)

Figure 2

Figure 2. (a) CV curves of the H2O-vapor-exposed battery recorded before (blue) and after (red) annealing in an Ar atmosphere at a temperature of 150 °C and pressure of 1 atm for 60 min. The measurements were conducted after the samples had been cooled down to ambient temperature. The voltage was swept in the range from 3.0 to 4.3 V vs Li/Li+ at a rate of 1 mV·s–1. (b) Impedance spectra recorded at voltages of 3.8 and 4.2 V vs Li+/Li for the H2O-vapor-exposed battery after annealing. The numbers denote the frequencies. (c) Dependence of the interface resistance on the annealing time obtained at a voltage of 4.2 V vs Li+/Li and an annealing temperature of 150 °C. The interface resistances were measured after cooling the sample to room temperature (27 °C). (d) Annealing temperature dependences of the interface resistance obtained at a voltage 4.2 V vs Li+/Li and annealing times of 60 and 120 min.

Reduction of the Interface Resistance by Battery Annealing

To our surprise, the annealing of the battery drastically reduced the interface resistance and restored the original battery performance. We annealed a sample after depositing Li3PO4 and an Li negative electrode. Figure 2a shows the CV curve of the H2O-vapor-exposed battery recorded after annealing at 150 °C in an Ar atmosphere for 60 min (hereafter called the battery-annealed sample). The peak current at the 20th cycle (∼3.9 V vs Li/Li+) is more than 90% of that of the in vacuo battery with the same electrode thickness. Moreover, the corresponding interface resistance is 10.3 Ω·cm2 (Figure 3b), which is comparable to the value obtained for the in vacuo battery (10.9 Ω·cm2). Furthermore, the cyclic battery performance was comparable to that of the in vacuo battery (Figure S5). The air-exposed batteries also showed recovery from 200 to 21.8 Ω·cm2 (Figure S6).

Figure 3

Figure 3. (a) X-ray crystal truncation rod (CTR) scattering profiles of the pristine (in vacuo, black), H2O-vapor-exposed (blue), and annealed (red) battery samples. The spectra were recorded after depositing Li3PO4 layers on the LiCoO2 surfaces. An Li thin film was deposited only on the Li3PO4 layer of the annealed battery sample. The thicknesses of the LiCoO2, Li3PO4, and Li thin films were 5 nm, 500 nm, and 1 μm, respectively. (b) Electron density profiles of the in vacuo, H2O-vapor-exposed, and annealed battery samples derived from the results of structural analyses of the corresponding CTR profiles. Owing to the existence of the antiphase domain, the electron density of the Li layer appears to be larger than the actual electron density of Li.

The degree of battery performance recovery depends on the annealing duration and temperature. As the time of annealing at 150 °C increases from 15 to 60 min, the interface resistance decreases, whereas it remains constant beyond 60 min (Figure 2c). In addition, we annealed the H2O-vapor-exposed samples for 60 and 120 min at various temperatures (Figure 2d). The interface resistance is lower at higher annealing temperatures and is saturated beyond an annealing duration of 60 min at each temperature. These results suggest that multiple thermodynamic factors play important roles in the recovery process.

Dependence of Performance Recovery on Stacking Conditions

To reveal the recovery mechanism, we annealed the sample with two different structures. First, an H2O-vapor-exposed LiCoO2 thin film was annealed without depositing Li3PO4 or Li metal. The LiCoO2 surface was annealed immediately after exposure (Figure S7a) at 200 °C for 60 min in a vacuum (<1 × 10–5 Pa) to desorb H2O molecules. After cooling to room temperature, Li3PO4 and Li electrode layers were deposited to form a battery. Although heating was performed at a temperature significantly exceeding the evaporation temperature of water in a vacuum, we observed no reduction in the interface resistance, and the solid-electrolyte–electrode interface remained degraded (Figure S7b,c).
The second method involved annealing the sample after Li3PO4 deposition, similarly to the procedure used in earlier studies (Figure S7b). (8,9) The LiCoO2 surface was exposed to H2O followed by the Li3PO4 deposition. Subsequently, the sample was annealed at 150 °C in an Ar atmosphere (1 atm) for 60 min. After the sample had been cooled to room temperature, an Li negative electrode was deposited to form a battery. The results showed that the CV peak current considerably increased, whereas the interface resistance decreased to 125 Ω·cm2 (Figure S7e,f). The latter was still an order of magnitude larger than the interface resistances of the in vacuo batteries. These investigations indicate that the battery structure is crucial for recovery.

Structural Analysis of the H2O-Vapor-Exposed Battery Interface

Next, the atomic structures of the H2O-vapor-exposed LiCoO2 film after the battery annealing were investigated. The obtained X-ray CTR scattering data indicate that the CoO2 layers near the LiCoO2 surface are robust. Figure 3a shows the diffraction patterns in which the horizontal axis represents the Miller index L in the direction normal to the thin-film surface. For all three battery samples (in vacuo, H2O-vapor-exposed, and annealed), the LiCoO2 003n (n = integer) Bragg peaks are located at the same positions, and thickness fringes are observed. These results suggest that the crystallinity degree of interface LiCoO2 remains high even after H2O vapor exposure and annealing treatment and that no secondary phase layer is formed in the LiCoO2 layers. The electron density profiles derived from the CTR scattering data contain sharp peaks even on the surfaces (Li3PO4–LiCoO2 interfaces) of the three different battery types (Figure 3b). Hence, the CoO2 layers remained intact after reaction with the H2O species. These results indicate that H critically contributes to the degradation and annealing recovery.

Discussion

Click to copy section linkSection link copied!

This section suggests a plausible mechanism of the interface resistance recovery proposed by density functional theory (DFT) calculations. Previous studies of H2O-vapor-exposed LiCoO2 surfaces using X-ray photoemission and absorption spectroscopy techniques (22,23) suggest that the formation of Li vacancies and incorporation of protons from dissociated H2O into the LiCoO2 crystal produce H-incorporated LiCoO2 species. (24) Furthermore, the X-ray CTR scattering data confirm the intactness of the layered structures (Figure 3b). Thus, the battery recovery process should involve filling the Li vacancies with Li atoms and removing protons from the LiCoO2 lattice.
The first-principles calculations predict that protons are spontaneously removed from LiCoO2 because of the differences between the electrochemical potentials of the protons in the LiCoO2 and Li structures. In other words, an all-solid-state proton battery (25) is spontaneously formed after Li electrode deposition. Thus, the proton battery self-discharges during the annealing process. Considering the charge neutrality of LiCoO2, the following reaction should occur at the Li3PO4–LiCoO2 interface:
(1)
Deintercalated protons migrate through the Li3PO4 electrolyte to the Li negative electrode side. Indeed, proton conductivity has previously been reported for the H3PO4 structure, (26) which has the same anionic structure as Li3PO4. (27) Inversely, Li ions migrate from the Li negative electrode through the Li3PO4 electrolyte layer to fill the vacant sites in the Li1–xCoO2 structure.
To verify this hypothesis, we performed first-principles calculations for various types of intrinsic defects and a proton in the LiCoO2 lattice. The obtained defect formation energies confirmed that Li vacancies were dominant among Li, Co, and O vacancies in the bulk LiCoO2. This result is consistent with the data reported previously. (28) Therefore, it is reasonable to consider the formation of Li vacancies upon H2O vapor exposure. When these vacancies are present in the crystal structure, EF is located at the top of the valence band (EF ≈ 0.0 eV), which matches the results of previous theoretical (29) and experimental (30) studies. Hence, Li1–xCoO2 represents a heavily doped p-type semiconductor. (31)
Next, we discuss the incorporation of protons into the LiCoO2 lattice. First, we considered a proton located at the Li vacancy site; however, the energy of this structure was higher than that of the interstitial site (Figure S8). Using the same method, we calculated the energies of several configurations containing interstitial protons in the LiCoO2 lattice (Figure S9). Their optimized structures depicted in Figure 4a indicate that the interstitial proton is strongly bonded to the O atoms in LiCoO2. (32) The measured O–H bond length is approximately 0.99 Å (Figure 4a) with a proton donating a charge of ∼0.5e. This value is close to the O–H bond length in H2O gas molecules (0.958 Å). (33)

Figure 4

Figure 4. (a) Left: optimized crystal structure of the bulk LiCoO2 unit cell with an interstitial H atom. Right: schematic of H–O bonding at the most stable site. The adjacent Li atom is shifted from the O and H atoms. The structures were visualized using the VESTA software package. (33) (b) Formation energies of the neutral and charged H interstitials in the bulk LiCoO2 structures plotted as functions of the Fermi energy (EF) with respect to the valence-band maximum. The H partial pressure was set to pH2 ≈ 10–9 atm. The kinks correspond to the (+/0) charge transition levels. (c) Energy diagram constructed for the diffusion of an interstitial H from the most stable site to the closest neighboring stable site. The activation energy (Eact) of the proton diffusion process was estimated by considering the zero-point energies of the initial and transition states. (d) Diffusion coefficients estimated for an interstitial H atom in the LiCoO2 structure at different annealing temperatures.

Figure 4b shows the formation energy of proton interstitials as a function of EF at different temperatures (T) and the low H partial pressure pH2 ≈10–4 Pa. The calculation results indicate that H0 species are unstable, whereas protons are formed when EF is near the top of the valence band. This result is consistent with the findings of a previous study on H defects in the p-type GaN, which stated that H solubility is significantly higher in p-type semiconductors. (31,34) Note that near the annealing temperature T ∼ 400 K, the formation energy of protons is shifted by approximately 0.9 eV from 0 K, indicating a decrease in stability with increasing temperature. These results are consistent with the removal of interstitial protons during annealing.
On the negative electrode side, there is a possibility of LiH formation due to the reaction of protons and Li. (35) The calculated Gibbs energies of the formation of LiH and proton interstitials in LiCoO2 are −0.765 and −0.153 eV, respectively. Therefore, protons are thermodynamically more stable in Li than in the LiCoO2 structure, leading to the spontaneous recovery of LiCoO2 after the deposition of the Li negative electrode. Furthermore, annealing the battery structure promotes proton migration and, therefore, the battery recovery process. In this work, we estimated the activation energy (Ea) of the proton migration in the LiCoO2 lattice, assuming that proton hopping involved only jumps from stable sites to the neighboring stable sites (Figure 4a). Figure 4c shows the energy barrier of the proton diffusion process. The Ea value estimated from the difference between the lowest energy and that of the saddle point is 1.41 eV at 0 K. Its magnitude is decreased to 1.278 eV when the zero-point energies of the initial state (0.333 eV) and transition state (0.201 eV) are considered (Figure 4d).
We can estimate the temperature at which the protons are sufficiently mobile from the condition that the calculated hopping rate (see the Methods section) is higher than 1 Hz. The calculated result is 495 K, which is comparable to the maximal experimental annealing temperature of 433 K. The diffusion coefficient of protons in the LiCoO2 structure increases with temperature (Figure 4d), and its value was estimated to be ∼10–17 cm2 s–1 at 433 K.
To validate the proposed recovery mechanism of hydrogen deintercalation, we measured the amount of protons using negative-ion time-of-flight secondary ion mass spectroscopy (ToF-SIMS). We tracked the ion signals of OH, Li, Al, P, and Co in the battery cell. The amount of protons was evaluated based on the amount of OH. First, we confirmed the hydrogen background level in the clean interface of Li3PO4 and LiCoO2 (Figure 5a). Before exposure to H2O, the ratio of the intensity of OH, hydrogen species, to the Co intensity of the LiCoO2 layer is 0.29 at a point at the center of the LiCoO2 layer. In contrast, the magnitude of the intensities of Co and OH is reversed after H2O exposure, and the ratio increases to 1.54 (Figure 5b). The exposed H2O-vapor-derived H diffuses into LiCoO2. Finally, after Li deposition and battery annealing, the intensity ratio is reduced to 0.56 (Figure 5c). This finding is consistent with the proton deintercalation results from LiCoO2 discussed in relation to the DFT calculations. Battery annealing was found to contribute to the normalization of the composition in the positive electrode.

Figure 5

Figure 5. (a–c) Negative ion time-of-flight secondary ion mass spectroscopy spectra for OH, Al, P, and Co. Figure S10 presents the data for Li and H. The green and blue regions represent Li3PO4 and LiCoO2, respectively. The signals of Al, P, and Co were used to identify the interfaces. (a) Sample without H2O exposure, (b) H2O-vapor-exposed battery, and (c) battery annealed after H2O vapor exposure. An Li layer was deposited on the annealed sample only. (d) Schematic illustration of the battery performance deterioration with a large interface resistance caused by water adsorption and recovery of the low interface resistance via battery annealing.

Conclusions

Click to copy section linkSection link copied!

In this study, the recovery of low electrolyte–electrode interface resistance was achieved by annealing the entire battery. Among the gas species present in air, only H2O degraded the interface between the amorphous Li3PO4 solid electrolyte and (001)-oriented LiCoO2 positive electrode, considerably increasing its resistance. The influence of protons on the properties of the solid-electrolyte–positive electrode interface was significant; however, during the fabrication of the battery structure, protons spontaneously migrated from LiCoO2 to the Li electrode through the Li3PO4 electrolyte (Figure 5d). The formation of a proton battery successfully recovered the initial low-resistance interface. The elucidation of the interfacial microscopic processes can help widen the application range of solid-state batteries.

Methods

Click to copy section linkSection link copied!

Fabrication of the Gas-Exposed Thin-Film Batteries

Thin-film deposition and sample transfer were performed inside an all-in-vacuum (in vacuo) system (13) except for the intentional gas exposures. An Au current collector (thickness: 100 nm) was deposited on an Al2O3 (0001) single crystal substrate by direct-current magnetron sputtering (power: 30 W, Ar gas partial pressure: 0.5 Pa) and annealed in a vacuum (<1 × 10–5 Pa) at a substrate temperature (Ts) of 600 °C. An LiCoO2 positive electrode (thickness: 40 nm) was deposited by radio frequency (RF) magnetron sputtering (power: 150 W, Ts = 600 °C, and Ar partial pressure: 0.5 Pa) using an Li1.2CoO2+δ target (Toshima Manufacturing Co. Ltd., Japan) followed by postdeposition annealing (Ts = 600 °C; O2 partial pressure: ∼9 Pa). The LiCoO2 surface was exposed to air or pure gas molecules (N2, CO2, O2, H2, and H2O) inside a high-vacuum gas-exposure chamber (<5 × 10–5 Pa). Table 2 lists the utilized exposure conditions. After the exposure, the chamber was re-evacuated to the vacuum and transferred back to the thin-film deposition chamber. An Li3PO4 solid electrolyte layer (thickness: 1 μm) was deposited by RF magnetron sputtering (power: 100 W; Ar partial pressure: 0.15 Pa). An Li metal negative electrode (thickness: 1 μm) was deposited by thermal evaporation. Both the Li3PO4 and Li layers were deposited on the unheated surfaces. Figure S1 shows the shapes and sizes of the prepared samples.
Table 2. Various Gas Species Exposed under Different Conditions
 conditions
gas speciespressure (Pa)exposure time (min)
N2∼1 × 10530
O2∼1 × 10530
CO2∼1 × 10530
H2∼1 × 10330
H2O∼2 × 10230
Air∼1 × 10530

Characterization of Thin Films and Battery Performance Evaluation

Electrochemical measurements were performed at room temperature (27 °C) inside an Ar-filled glove box (O2 < 0.5 ppm; H2O < 0.25 ppm) using a frequency response analyzer (Biologic SP-150). The detailed interface resistance determination process is described in Figure S2. The thicknesses of the obtained thin films were determined using a Stylus profiler (Veeco, Dek-Tak 150).

X-ray CTR Scattering Analysis

X-ray CTR scattering analysis (36) was performed to determine the stacking structure of the LiCoO2 layers. For the atomic-scale analysis, we used LiCoO2 epitaxial thin films on an Al2O3(0001) single crystalline substrate prepared by pulsed laser deposition with a KrF excimer laser (see the Supplementary Information for the details of the deposition process). The battery-annealed sample was placed in a vacuum sample cell (<1 × 10–4 Pa) during the measurements to avoid the deterioration of the Li layer. Other samples without Li layers were placed in a sample cell with He gas flow during the measurements. The CTR measurements were performed at the BL13XU beamline of the SPring-8 synchrotron radiation facility and BL-3A beamline of the Photon Factory, KEK. The incident X-ray energy was 11 keV, and the scattered X-rays were detected using a two-dimensional detector (PILATUS-100 K). The antiphase domain was considered in the analysis of the electron density profile. (37)

DFT Calculation

To gain insight into the H defects of the LiCoO2 structure, we performed first-principles calculations using the projector augmented-wave method (38) with a plane-wave cutoff energy of 520 eV and the optPBE-vdW functional (39−41) implemented in the VASP software package. (42) We used a gamma-centered 5 × 5 × 5 k-points set for Brillouin zone integration. The electron correlations in the 3d orbitals of LiCoO2 Co atoms were considered through onsite Hubbard correction (DFT + U, U = 3.32 eV). (43) The calculations produced a bandgap of approximately 2.1 eV for pristine LiCoO2 and an open-circuit voltage (VOC) of approximately 4.07 V. The stability of a point defect X in a charge state q was determined from the formation energy via the following formula: (44)
(2)
where ET(Xq) and ET(LCO) are the DFT total energies of the defective system and bulk LiCoO2 supercell, respectively; ni is the number of atomic species i with chemical potential μi that are removed from (ni < 0) or added to (ni > 0) the supercell; EVBM is the energy corresponding to the valence-band maximum (VBM); EF is the Fermi energy; and Δcorrq accounts for the spurious interactions between the charged defects and periodic supercell images. (45)EF can be derived by equating the sums of the concentrations of all negative and positive defects, which requires estimation of the formation energies of all intrinsic defects in the bulk LiCoO2 structure.
We obtained the total energies by performing DFT simulations on a 108-atom 3 × 3 × 1 LiCoO2 supercell. To determine the Li chemical potential at the positive electrode, we set the negative electrode as a reference and assumed μLi = – eVOC = – 4.07 eV for all Ef(Xq). The chemical potential of H defects was calculated as , (46) where the first term corresponds to half of the total DFT energy of an H2 molecule in the gas phase at 0 K, and the last two terms account for the effects of temperature and the hydrogen partial pressure. The migration barrier was estimated by computing the total energy curve of a fixed H atom at different positions along the path connecting the initial and final states while allowing the surrounding atoms to relax. The diffusion coefficient was calculated as follows: (47), where . Ea is the activation energy, n is the number of sites, d is the dimension, and r is the jump length. The prefactor (attempt frequency) THz ∼1013 Hz was derived from the DFT simulation results using Vineyard’s method. (44) Here, viIS and viTS are the DFT vibrational frequencies of the initial and transition states, respectively. We calculated the annealing temperature by assuming a jump rate of 1 Hz (48) and using our computed prefactor.

Negative ion ToF-SIMS analysis

To confirm the hydrogen distribution across the thin-film battery, we analyzed the H2O-vapor-exposed Li3PO4–LiCoO2 interface by performing ToF-SIMS. A primary beam with 60 keV Bi3+ and a measured current of 0.06 Pa were used for the analysis. The sputtering was performed with 2 keV Cs+ ions on a 300 μm square area, and analysis was performed in a 100 μm square.

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.1c17945.

  • Fabrication of gas-exposed thin-film batteries, X-ray crystal truncation rod (CTR) scattering analysis, estimation of the interface resistance using an equivalent circuit, electrochemical properties of the pure gas-exposed batteries, impedance spectra of the H2O-vapor-exposed battery, parameters of the equivalent circuits used for the gas-exposed batteries, capacity retention, annealing recovery of the air-exposed battery, annealing processes and performance of the H2O-vapor-exposed battery, parameters of the equivalent circuits used for the gas-exposed batteries after the annealing process, optimized geometry of a proton occupying a Li vacancy site in the LiCoO2 structure, comparison of different LCO proton interstitial sites, and the time-of-flight secondary ion mass spectroscopy spectrum for H and Li species (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 Authors
  • Authors
    • Elvis F. Arguelles - Department of Materials Engineering, The University of Tokyo, Tokyo 113-8656, Japan
    • Tetsuroh Shirasawa - National Metrology Institute of Japan, National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba, Ibaraki 305-8565, JapanOrcidhttps://orcid.org/0000-0001-5519-6977
    • Shusuke Kasamatsu - Faculty of Science, Yamagata University, Yamagata 990-8560, JapanOrcidhttps://orcid.org/0000-0002-4368-7995
    • Koji Shimizu - Department of Materials Engineering, The University of Tokyo, Tokyo 113-8656, Japan
    • Kazunori Nishio - School of Materials and Chemical Technology, Tokyo Institute of Technology, Meguro, Tokyo 152-8552, JapanOrcidhttps://orcid.org/0000-0002-8201-358X
    • Yuki Watanabe - School of Materials and Chemical Technology, Tokyo Institute of Technology, Meguro, Tokyo 152-8552, JapanOrcidhttps://orcid.org/0000-0003-2589-1105
    • Yusuke Kubota - Tokyo Electron Technology Solutions Limited, 650 Mitsuzawa, Hosaka−cho, Nirasaki, Yamanashi 407-0192, Japan
    • Ryota Shimizu - School of Materials and Chemical Technology, Tokyo Institute of Technology, Meguro, Tokyo 152-8552, JapanPRESTO, Japan Science and Technology Agency, Kawaguchi, Saitama 332-0012, JapanOrcidhttps://orcid.org/0000-0001-9600-7255
    • Satoshi Watanabe - Department of Materials Engineering, The University of Tokyo, Tokyo 113-8656, JapanOrcidhttps://orcid.org/0000-0002-8069-6938
  • Author Contributions

    S.Ko., Y.K., R.S., and T.H. designed the experiments. S.Ko. contributed to the fabrication and characterization of the thin films and thin-film batteries. E.F.A. contributed to the first-principles calculations of the hydrogen defect in the thin-film materials. T.S. contributed to the measurement and structural analysis of synchrotron X-ray CTR scattering. S.Ka., K.S., and S.W. supported the first-principles calculations and discussed the calculation settings. Y.W. and K.N. supported the thin-film fabrications. S.Ko. and T.H. wrote the manuscript with help from all of the co-authors.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

We thank Kazuhiko Kikuchi and Haruo Nagasawa (Pascal Co.) for experimental support. We thank Dr. Daiki Katsube (Nagaoka University of Technology) for technical advice on H2O vapor purification. We thank Akira Genseki at the Center for Advanced Materials Analysis, Tokyo Institute of Technology, for assistance with the ToF-SIMS measurements. We also thank Dr. Toshiya Saito (Toyota Corporation) and Prof. Clare P. Grey (University of Cambridge) for fruitful discussions. E.F.A. acknowledges funding from JSPS KAKENHI (grant no. 19K15397), Japan. R.S. acknowledges funding from JSPS KAKENHI (grant no. 17H05216) and JST–PRESTO (grant no. JPMJPR17N6), Japan. T.H. acknowledges funding from JSPS KAKENHI (grant nos. JP18H03876 and JP18H05514) and the JST–CREST (grant no. JPMJCR1523) program. The X-ray CTR experiments were conducted at the SPring–8 synchrotron radiation facility (proposal Nos. 2018A1256 and 2018B1286) and Photon Factory, KEK (PF–PAC No. 2019G056). We would like to thank Editage (www.editage.jp) for English language editing.

References

Click to copy section linkSection link copied!

This article references 48 other publications.

  1. 1
    Tarascon, J.-M.; Armand, M. Issues and Challenges Facing Rechargeable Lithium Batteries. Nature 2001, 414, 359367,  DOI: 10.1038/35104644
  2. 2
    Manthiram, A.; Yu, X.; Wang, S. Lithium Battery Chemistries Enabled by Solid-state Electrolytes. Nat. Rev. Mater. 2017, 2, 16103,  DOI: 10.1038/natrevmats.2016.103
  3. 3
    Wu, J.; Shen, L.; Zhang, Z.; Liu, G.; Wang, Z.; Zhou, D.; Wan, H.; Xu, X.; Yao, X. All-Solid-State Lithium Batteries with Sulfide Electrolytes and Oxide Cathodes. Electrochem. Energy Rev. 2021, 4, 101135,  DOI: 10.1007/s41918-020-00081-4
  4. 4
    Wu, J.; Sufu, L.; Han, F.; Yao, X.; Wang, C. Lithium/Sulfide All-Solid-State Batteries using Sulfide Electrolytes. Adv. Mater. 2020, 33, 2000751  DOI: 10.1002/adma.202000751
  5. 5
    Tan, D. H. S.; Chen, Y.; Yang, H.; Bao, W.; Sreenarayanan, B.; Doux, J.; Li, W.; Lu, B.; Ham, S.; Sayahpour, B.; Scharf, J.; Wu, E. A.; Deysher, G.; Han, H. E.; Hah, H. J.; Jeong, H.; Lee, J. B.; Chen, Z.; Meng, Y. S. Carbon-free high-loading silicon anodes enabled by sulfide solid electrolytes. Science 2021, 373, 14941499,  DOI: 10.1126/science.abg7217
  6. 6
    Yamada, H.; Tsunoe, D.; Shiraishi, S.; Isomichi, G. Reduced Grain Boundary Resistance by Surface Modification. J. Phys. Chem. C 2015, 119, 54125419,  DOI: 10.1021/jp510077z
  7. 7
    Gittleson, F. S.; el Gabaly, F. Non-faradaic Li+ Migration and Chemical Coordination Across Solid-state Battery Interfaces. Nano Lett. 2017, 17, 69746982,  DOI: 10.1021/acs.nanolett.7b03498
  8. 8
    Yada, C.; Ohmori, A.; Ide, K.; Yamasaki, H.; Kato, T.; Saito, T.; Sagane, F.; Iriyama, Y. Dielectric Modification of 5V-class Cathodes for High-voltage All-solid-state Lithium Batteries. Adv. Energy Mater. 2014, 4, 1301416  DOI: 10.1002/aenm.201301416
  9. 9
    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
  10. 10
    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
  11. 11
    Iriyama, Y.; Kako, T.; Yada, C.; Abe, T.; Ogumi, Z. Reduction of Charge Transfer Resistance at the Lithium Phosphorus Oxynitride/Lithium Cobalt Oxide Interface by Thermal Treatment. J. Power Sources 2005, 146, 745748,  DOI: 10.1016/j.jpowsour.2005.03.073
  12. 12
    Iriyama, Y.; Kako, T.; Yada, C.; Abe, T.; Ogumi, Z. Charge Transfer Reaction at the Lithium Phosphorus Oxynitride Glass Electrolyte/Lithium Cobalt Oxide Thin Film Interface. Solid State Ionics 2005, 176, 23712376,  DOI: 10.1016/j.ssi.2005.02.025
  13. 13
    Faenza, N. V.; Bruce, L.; Lebens-Higgins, Z. W.; Plitz, I.; Pereira, N.; Piper, L. F.; Amatucci, G. G. Growth of Ambient Induced Surface Impurity Species on Layered Positive Electrode Materials and Impact on Electrochemical Performance. J. Electrochem. Soc. 2017, 164, A3727A3741,  DOI: 10.1149/2.0921714jes
  14. 14
    Sicklinger, J.; Metzger, M.; Beyer, H.; Pritzl, D.; Gasteiger, H. A. Ambient Storage Derived Surface Contamination of NCM811 and NCM111: Performance Implications and Mitigation Strategies. J. Electrochem. Soc. 2019, 166, A2322A2335,  DOI: 10.1149/2.0011912jes
  15. 15
    Takada, K.; Ohta, N.; Zhang, L.; Fukuda, K.; Sakaguchi, I.; Ma, R.; Osada, M.; Sasaki, T. Interfacial modification for high-power solid-state lithium batteries. Solid State Ionics 2008, 179, 13331337,  DOI: 10.1016/j.ssi.2008.02.017
  16. 16
    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
  17. 17
    Haruta, M.; Shiraki, S.; Suzuki, T.; Kumatani, A.; Ohsawa, T.; Takagi, Y.; Shimizu, R.; Hitosugi, T. Negligible “Negative Space-charge Layer Effects” at Oxide-electrolyte/Electrode Interfaces of Thin-film Batteries. Nano Lett. 2015, 15, 14981502,  DOI: 10.1021/nl5035896
  18. 18
    Kuwata, N.; Iwagami, N.; Tanji, Y.; Matsuda, Y.; Kawamura, J. Characterization of Thin-film Lithium Batteries with Stable Thin-film Li3PO4 Solid Electrolytes Fabricated by ArF Excimer Laser Deposition. J. Electrochem. Soc. 2010, 157, A521A527,  DOI: 10.1149/1.3306339
  19. 19
    Garcia, B.; Farcy, J.; Pereira-Ramos, J. P.; Baffier, N. Electrochemical Properties of Low Temperature Crystallized LiCoO2. J. Electrochem. Soc. 1997, 114, 11791184
  20. 20
    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
  21. 21
    Zou, L.; He, Y.; Liu, Z.; Jia, H.; Zhu, J.; Zheng, J.; Wang, G.; Li, X.; Xiao, J.; Liu, J.; Zhang, J. G.; Chen, G.; Wang, C. Unlocking the Passivation Nature of the Cathode–Air Interfacial Reactions in Lithium Ion Batteries. Nat. Commun. 2020, 11, 3204,  DOI: 10.1038/s41467-020-17050-6
  22. 22
    Motzko, M.; Carrillo Solano, M. A.; Jaegermann, W.; Hausbrand, R. Photoemission Study on the Interaction Between LiCoO2 Thin Films and Adsorbed Water. J. Phys. Chem. C 2015, 119, 2340723412,  DOI: 10.1021/acs.jpcc.5b05793
  23. 23
    Cherkashinin, G.; Jaegermann, W. Dissociative Adsorption of H2O on LiCoO2(00l) Surfaces: Co Reduction Induced by Electron Transfer from Intrinsic Defects. J. Chem. Phys. 2016, 144, 184706,  DOI: 10.1063/1.4948610
  24. 24
    Larcher, D.; Palacín, M. R.; Amatucci, G. G.; Tarascon, J. M. Electrochemically Active LiCoO2 and LiNiO2 Made by Cationic Exchange Under Hydrothermal Conditions. J. Electrochem. Soc. 1997, 144, 408417,  DOI: 10.1149/1.1837424
  25. 25
    Mishra, K.; Pundir, S. S.; Rai, D. K. All-solid-state Proton Battery Using Gel Polymer Electrolyte. AIP Conf. Proc. 2014, 1591, 633,  DOI: 10.1063/1.4872700
  26. 26
    Smith, J. P.; Brown, W. E.; Lehr, J. R. Structure of Crystalline Phosphoric Acid. J. Am. Chem. Soc. 1955, 77, 27282730,  DOI: 10.1021/ja01615a013
  27. 27
    Dippel, T.; Kreuer, K.; Lassegues, J.; Rodriguez, D. Proton Conductivity in Fused Phosphoric Acid; A 1H/31P PFG-NMR and QNS Study. Solid State Ionics 1993, 61, 4146,  DOI: 10.1016/0167-2738(93)90332-W
  28. 28
    Swift, M. W.; Qi, Y. First-principles Prediction of Potentials and Space-charge Layers in All-solid-state Batteries. Phys. Rev. Lett. 2019, 122, 167701  DOI: 10.1103/PhysRevLett.122.167701
  29. 29
    Neugebauer, J.; Van de Walle, C. G. Hydrogen in GaN: Novel Aspects of a Common Impurity. Phys. Rev. Lett. 1995, 75, 44524455,  DOI: 10.1103/PhysRevLett.75.4452
  30. 30
    Fingerle, M.; Buchheit, R.; Sicolo, S.; Albe, K.; Hausbrand, R. Reaction and Space Charge Layer Formation at the LiCoO2–LiPON Interface: Insights on Defect Formation and Ion Energy Level Alignment by a Combined Surface Science–Simulation Approach. Chem. Mater. 2017, 29, 76757685,  DOI: 10.1021/acs.chemmater.7b00890
  31. 31
    Tukamoto, H.; West, A. R. Electronic Conductivity of LiCoO2 and Its Enhancement by Magnesium Doping. J. Electrochem. Soc. 1997, 144, 31643168,  DOI: 10.1149/1.1837976
  32. 32
    Momma, K.; Izumi, F. VESTA 3 for Three-dimensional Visualization of Crystal, Volumetric and Morphology Data. J. Appl. Crystallogr. 2011, 44, 12721276,  DOI: 10.1107/S0021889811038970
  33. 33
    Johnson, III, R. D. NIST101. Computational Chemistry Comparison and Benchmark Database, https://cccbdb.nist.gov/.
  34. 34
    Van de Walle, C. G.; Neugebauer, J. Universal Alignment of Hydrogen Levels in Semiconductors, Insulators and Solutions. Nature 2003, 423, 626628,  DOI: 10.1038/nature01665
  35. 35
    Xu, G.; Li, J.; Wang, C.; du, X.; Lu, D.; Xie, B.; Wang, X.; Lu, C.; Liu, H.; Dong, S.; Cui, G.; Chen, L. The Formation/Decomposition Equilibrium of LiH and Its Contribution on Anode Failure in Practical Lithium Metal Batteries. Angew. Chem., Int. Ed. 2021, 60, 77707776,  DOI: 10.1002/anie.202013812
  36. 36
    Wakabayashi, Y.; Shirasawa, T.; Voegeli, W.; Takahashi, T. Observation of Structure of Surfaces and Interfaces by Synchrotron X-ray Diffraction: Atomic-scale Imaging and Time-resolved Measurements. J. Phys. Soc. Jpn. 2018, 87, 061010  DOI: 10.7566/JPSJ.87.061010
  37. 37
    Zheng, S. J.; Fisher, C. A. J.; Hitosugi, T.; Kumatani, A.; Shiraki, S.; Ikuhara, Y. H.; Kuwabara, A.; Moriwake, H.; Oki, H.; Ikuhara, Y. Antiphase Inversion Domains in Lithium Cobaltite Thin Films Deposited on Single-crystal Sapphire Substrates. Acta Mater. 2013, 61, 76717678,  DOI: 10.1016/j.actamat.2013.09.004
  38. 38
    Blöchl, P. E. Projector Augmented-wave Method. Phys. Rev. B 1994, 50, 1795317979,  DOI: 10.1103/PhysRevB.50.17953
  39. 39
    Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865,  DOI: 10.1103/PhysRevLett.77.3865
  40. 40
    Klimeš, J.; Bowler, D. R.; Michaelides, A. Chemical Accuracy for the van der Waals Density Functional. J. Phys. Condens. Matter. 2010, 22, 022201  DOI: 10.1088/0953-8984/22/2/022201
  41. 41
    Klimeš, J.; Bowler, D. R.; Van der Michaelides, A. Waals Density Functionals Applied to Solids. Phys. Rev. B 2011, 83, 195131  DOI: 10.1103/PhysRevB.83.195131
  42. 42
    Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-energy Calculations Using a Plane-wave Basis Set. Phys. Rev. B 1996, 54, 1116911186,  DOI: 10.1103/PhysRevB.54.11169
  43. 43
    Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.; Sutton, A. P. Electron-energy-loss Spectra and the Structural Stability of Nickel Oxide: An LSDA+ U Study. Phys. Rev. B 1998, 57, 15051509,  DOI: 10.1103/PhysRevB.57.1505
  44. 44
    Freysoldt, C.; Neugebauer, J.; Van de Walle, C. G. Fully Ab Initio Finite-size Corrections for Charged-defect Supercell Calculations. Phys. Rev. Lett. 2009, 102, 016402  DOI: 10.1103/PhysRevLett.102.016402
  45. 45
    Imamoğlu, A. Cavity QED Based on Collective Magnetic Dipole Coupling: Spin Ensembles as Hybrid Two-level Systems. Phys. Rev. Lett. 2009, 102, 083602  DOI: 10.1103/PhysRevLett.102.083602
  46. 46
    Reuter, K.; Scheffler, M. Composition, Structure, and Stability of RuO2 (110) as a Function of Oxygen Pressure. Phys. Rev. B 2001, 65, 035406  DOI: 10.1103/PhysRevB.65.035406
  47. 47
    Koettgen, J.; Zacherle, T.; Grieshammer, S.; Martin, M. Ab Initio Calculation of the Attempt Frequency of Oxygen Diffusion in Pure and Samarium Doped Ceria. Phys. Chem. Chem. Phys. 2017, 19, 99579973,  DOI: 10.1039/c6cp04802a
  48. 48
    Kyrtsos, A.; Matsubara, M.; Bellotti, E. Migration Mechanisms and Diffusion Barriers of Carbon and Native Point Defects in GaN. Phys. Rev. B 2016, 93, 245201  DOI: 10.1103/PhysRevB.93.245201

Cited By

Click to copy section linkSection link copied!

This article is cited by 8 publications.

  1. Adi Aizat Razali, Siti Nurshahira Norazli, Wei Siang Sum, Siew Yean Yeo, Andrea Dolfi, Geetha Srinivasan. State-of-the-Art of Solid-State Electrolytes on the Road Map of Solid-State Lithium Metal Batteries for E-Mobility. ACS Sustainable Chemistry & Engineering 2023, 11 (21) , 7927-7964. https://doi.org/10.1021/acssuschemeng.3c00057
  2. Shigeru Kobayashi, Kazunori Nishio, Markus Wilde, Katsuyuki Fukutani, Ryota Shimizu, Taro Hitosugi. Protons Inside the LiCoO2 Electrode Largely Increase Electrolyte–Electrode Interface Resistance in All-Solid-State Li Batteries. The Journal of Physical Chemistry C 2023, 127 (9) , 4684-4688. https://doi.org/10.1021/acs.jpcc.2c06620
  3. Hoai Khang Tran, Beta Thi Truong, Bo-Rong Zhang, Rajan Jose, Jeng-Kuei Chang, Chun-Chen Yang. Sandwich-Structured Composite Polymer Electrolyte Based on PVDF-HFP/PPC/Al-Doped LLZO for High-Voltage Solid-State Lithium Batteries. ACS Applied Energy Materials 2023, 6 (3) , 1475-1487. https://doi.org/10.1021/acsaem.2c03363
  4. Jun Deng, Kazunori Nishio, Satoru Ichinokura, Yuki Watanabe, Kurei Edamura, Ryo Nakayama, Ryota Shimizu, Toru Hirahara, Taro Hitosugi. Modifying the Interface between the Solvated Ionic Liquid Electrolyte and Positive Electrode to Boost Lithium-Ion Battery Performance. ACS Applied Energy Materials 2022, 5 (9) , 10891-10896. https://doi.org/10.1021/acsaem.2c01533
  5. Kazunori Nishio, Daisuke Imazeki, Kosuke Kurushima, Yuki Takeda, Kurei Edamura, Ryo Nakayama, Ryota Shimizu, Taro Hitosugi. Immense Reduction in Interfacial Resistance between Sulfide Electrolyte and Positive Electrode. ACS Applied Materials & Interfaces 2022, 14 (30) , 34620-34626. https://doi.org/10.1021/acsami.2c05896
  6. Tsuyoshi Ohnishi, Kazunori Takada. Sputter-Deposited Amorphous Li3PO4 Solid Electrolyte Films. ACS Omega 2022, 7 (24) , 21199-21206. https://doi.org/10.1021/acsomega.2c02104
  7. Taro HITOSUGI, Kazunori NISHIO. Quantitative Understanding of the Resistance at the Solid Electrolyte and Electrode Interfaces. Denki Kagaku 2022, 90 (4) , 320-325. https://doi.org/10.5796/denkikagaku.22-FE0028
  8. Yosef Nikodimos, Wei-Nien Su, Bereket Woldegbreal Taklu, Semaw Kebede Merso, Teklay Mezgebe Hagos, Chen-Jui Huang, Haylay Ghidey Redda, Chia-Hsin Wang, She-Huang Wu, Chun-Chen Yang, Bing Joe Hwang. Resolving anodic and cathodic interface-incompatibility in solid-state lithium metal battery via interface infiltration of designed liquid electrolytes. Journal of Power Sources 2022, 535 , 231425. https://doi.org/10.1016/j.jpowsour.2022.231425
Open PDF

ACS Applied Materials & Interfaces

Cite this: ACS Appl. Mater. Interfaces 2022, 14, 2, 2703–2710
Click to copy citationCitation copied!
https://doi.org/10.1021/acsami.1c17945
Published January 6, 2022

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

CC-BY-NC-ND 4.0 .

Article Views

7698

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. Electrochemical characteristics of the (a, c, and e) in vacuo and (b, d, and f) air-exposed batteries. cyclic voltammetry (CV) curves of the (a) in vacuo and (b) air-exposed batteries. The voltage was swept at a rate of 1 mV·s–1 from 3.0 to 4.3 V vs Li+/Li. The displayed curves correspond to the 20th charge–discharge cycle. Impedance spectra of the (c) in vacuo and (d) air-exposed batteries recorded at voltages of 3.8 and 4.2 V vs Li+/Li. The numbers n in the plots indicate frequencies (10n Hz). Low-frequency regions of the impedance spectra recorded for the (e) in vacuo and (f) air-exposed batteries. The fitting data obtained for the equivalent circuit at 4.2 V vs Li+/Li data (Table S1) are shown as colored semicircles. The light-green semicircle denotes the resistance of Li ion conduction in the Li3PO4 solid electrolyte. The gray semicircle represents the Li3PO4–LiCoO2 interface resistance (10.9 Ω·cm2 for the in vacuo battery and 200 Ω·cm2 for the air-exposed battery).

    Figure 2

    Figure 2. (a) CV curves of the H2O-vapor-exposed battery recorded before (blue) and after (red) annealing in an Ar atmosphere at a temperature of 150 °C and pressure of 1 atm for 60 min. The measurements were conducted after the samples had been cooled down to ambient temperature. The voltage was swept in the range from 3.0 to 4.3 V vs Li/Li+ at a rate of 1 mV·s–1. (b) Impedance spectra recorded at voltages of 3.8 and 4.2 V vs Li+/Li for the H2O-vapor-exposed battery after annealing. The numbers denote the frequencies. (c) Dependence of the interface resistance on the annealing time obtained at a voltage of 4.2 V vs Li+/Li and an annealing temperature of 150 °C. The interface resistances were measured after cooling the sample to room temperature (27 °C). (d) Annealing temperature dependences of the interface resistance obtained at a voltage 4.2 V vs Li+/Li and annealing times of 60 and 120 min.

    Figure 3

    Figure 3. (a) X-ray crystal truncation rod (CTR) scattering profiles of the pristine (in vacuo, black), H2O-vapor-exposed (blue), and annealed (red) battery samples. The spectra were recorded after depositing Li3PO4 layers on the LiCoO2 surfaces. An Li thin film was deposited only on the Li3PO4 layer of the annealed battery sample. The thicknesses of the LiCoO2, Li3PO4, and Li thin films were 5 nm, 500 nm, and 1 μm, respectively. (b) Electron density profiles of the in vacuo, H2O-vapor-exposed, and annealed battery samples derived from the results of structural analyses of the corresponding CTR profiles. Owing to the existence of the antiphase domain, the electron density of the Li layer appears to be larger than the actual electron density of Li.

    Figure 4

    Figure 4. (a) Left: optimized crystal structure of the bulk LiCoO2 unit cell with an interstitial H atom. Right: schematic of H–O bonding at the most stable site. The adjacent Li atom is shifted from the O and H atoms. The structures were visualized using the VESTA software package. (33) (b) Formation energies of the neutral and charged H interstitials in the bulk LiCoO2 structures plotted as functions of the Fermi energy (EF) with respect to the valence-band maximum. The H partial pressure was set to pH2 ≈ 10–9 atm. The kinks correspond to the (+/0) charge transition levels. (c) Energy diagram constructed for the diffusion of an interstitial H from the most stable site to the closest neighboring stable site. The activation energy (Eact) of the proton diffusion process was estimated by considering the zero-point energies of the initial and transition states. (d) Diffusion coefficients estimated for an interstitial H atom in the LiCoO2 structure at different annealing temperatures.

    Figure 5

    Figure 5. (a–c) Negative ion time-of-flight secondary ion mass spectroscopy spectra for OH, Al, P, and Co. Figure S10 presents the data for Li and H. The green and blue regions represent Li3PO4 and LiCoO2, respectively. The signals of Al, P, and Co were used to identify the interfaces. (a) Sample without H2O exposure, (b) H2O-vapor-exposed battery, and (c) battery annealed after H2O vapor exposure. An Li layer was deposited on the annealed sample only. (d) Schematic illustration of the battery performance deterioration with a large interface resistance caused by water adsorption and recovery of the low interface resistance via battery annealing.

  • References


    This article references 48 other publications.

    1. 1
      Tarascon, J.-M.; Armand, M. Issues and Challenges Facing Rechargeable Lithium Batteries. Nature 2001, 414, 359367,  DOI: 10.1038/35104644
    2. 2
      Manthiram, A.; Yu, X.; Wang, S. Lithium Battery Chemistries Enabled by Solid-state Electrolytes. Nat. Rev. Mater. 2017, 2, 16103,  DOI: 10.1038/natrevmats.2016.103
    3. 3
      Wu, J.; Shen, L.; Zhang, Z.; Liu, G.; Wang, Z.; Zhou, D.; Wan, H.; Xu, X.; Yao, X. All-Solid-State Lithium Batteries with Sulfide Electrolytes and Oxide Cathodes. Electrochem. Energy Rev. 2021, 4, 101135,  DOI: 10.1007/s41918-020-00081-4
    4. 4
      Wu, J.; Sufu, L.; Han, F.; Yao, X.; Wang, C. Lithium/Sulfide All-Solid-State Batteries using Sulfide Electrolytes. Adv. Mater. 2020, 33, 2000751  DOI: 10.1002/adma.202000751
    5. 5
      Tan, D. H. S.; Chen, Y.; Yang, H.; Bao, W.; Sreenarayanan, B.; Doux, J.; Li, W.; Lu, B.; Ham, S.; Sayahpour, B.; Scharf, J.; Wu, E. A.; Deysher, G.; Han, H. E.; Hah, H. J.; Jeong, H.; Lee, J. B.; Chen, Z.; Meng, Y. S. Carbon-free high-loading silicon anodes enabled by sulfide solid electrolytes. Science 2021, 373, 14941499,  DOI: 10.1126/science.abg7217
    6. 6
      Yamada, H.; Tsunoe, D.; Shiraishi, S.; Isomichi, G. Reduced Grain Boundary Resistance by Surface Modification. J. Phys. Chem. C 2015, 119, 54125419,  DOI: 10.1021/jp510077z
    7. 7
      Gittleson, F. S.; el Gabaly, F. Non-faradaic Li+ Migration and Chemical Coordination Across Solid-state Battery Interfaces. Nano Lett. 2017, 17, 69746982,  DOI: 10.1021/acs.nanolett.7b03498
    8. 8
      Yada, C.; Ohmori, A.; Ide, K.; Yamasaki, H.; Kato, T.; Saito, T.; Sagane, F.; Iriyama, Y. Dielectric Modification of 5V-class Cathodes for High-voltage All-solid-state Lithium Batteries. Adv. Energy Mater. 2014, 4, 1301416  DOI: 10.1002/aenm.201301416
    9. 9
      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
    10. 10
      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
    11. 11
      Iriyama, Y.; Kako, T.; Yada, C.; Abe, T.; Ogumi, Z. Reduction of Charge Transfer Resistance at the Lithium Phosphorus Oxynitride/Lithium Cobalt Oxide Interface by Thermal Treatment. J. Power Sources 2005, 146, 745748,  DOI: 10.1016/j.jpowsour.2005.03.073
    12. 12
      Iriyama, Y.; Kako, T.; Yada, C.; Abe, T.; Ogumi, Z. Charge Transfer Reaction at the Lithium Phosphorus Oxynitride Glass Electrolyte/Lithium Cobalt Oxide Thin Film Interface. Solid State Ionics 2005, 176, 23712376,  DOI: 10.1016/j.ssi.2005.02.025
    13. 13
      Faenza, N. V.; Bruce, L.; Lebens-Higgins, Z. W.; Plitz, I.; Pereira, N.; Piper, L. F.; Amatucci, G. G. Growth of Ambient Induced Surface Impurity Species on Layered Positive Electrode Materials and Impact on Electrochemical Performance. J. Electrochem. Soc. 2017, 164, A3727A3741,  DOI: 10.1149/2.0921714jes
    14. 14
      Sicklinger, J.; Metzger, M.; Beyer, H.; Pritzl, D.; Gasteiger, H. A. Ambient Storage Derived Surface Contamination of NCM811 and NCM111: Performance Implications and Mitigation Strategies. J. Electrochem. Soc. 2019, 166, A2322A2335,  DOI: 10.1149/2.0011912jes
    15. 15
      Takada, K.; Ohta, N.; Zhang, L.; Fukuda, K.; Sakaguchi, I.; Ma, R.; Osada, M.; Sasaki, T. Interfacial modification for high-power solid-state lithium batteries. Solid State Ionics 2008, 179, 13331337,  DOI: 10.1016/j.ssi.2008.02.017
    16. 16
      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
    17. 17
      Haruta, M.; Shiraki, S.; Suzuki, T.; Kumatani, A.; Ohsawa, T.; Takagi, Y.; Shimizu, R.; Hitosugi, T. Negligible “Negative Space-charge Layer Effects” at Oxide-electrolyte/Electrode Interfaces of Thin-film Batteries. Nano Lett. 2015, 15, 14981502,  DOI: 10.1021/nl5035896
    18. 18
      Kuwata, N.; Iwagami, N.; Tanji, Y.; Matsuda, Y.; Kawamura, J. Characterization of Thin-film Lithium Batteries with Stable Thin-film Li3PO4 Solid Electrolytes Fabricated by ArF Excimer Laser Deposition. J. Electrochem. Soc. 2010, 157, A521A527,  DOI: 10.1149/1.3306339
    19. 19
      Garcia, B.; Farcy, J.; Pereira-Ramos, J. P.; Baffier, N. Electrochemical Properties of Low Temperature Crystallized LiCoO2. J. Electrochem. Soc. 1997, 114, 11791184
    20. 20
      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
    21. 21
      Zou, L.; He, Y.; Liu, Z.; Jia, H.; Zhu, J.; Zheng, J.; Wang, G.; Li, X.; Xiao, J.; Liu, J.; Zhang, J. G.; Chen, G.; Wang, C. Unlocking the Passivation Nature of the Cathode–Air Interfacial Reactions in Lithium Ion Batteries. Nat. Commun. 2020, 11, 3204,  DOI: 10.1038/s41467-020-17050-6
    22. 22
      Motzko, M.; Carrillo Solano, M. A.; Jaegermann, W.; Hausbrand, R. Photoemission Study on the Interaction Between LiCoO2 Thin Films and Adsorbed Water. J. Phys. Chem. C 2015, 119, 2340723412,  DOI: 10.1021/acs.jpcc.5b05793
    23. 23
      Cherkashinin, G.; Jaegermann, W. Dissociative Adsorption of H2O on LiCoO2(00l) Surfaces: Co Reduction Induced by Electron Transfer from Intrinsic Defects. J. Chem. Phys. 2016, 144, 184706,  DOI: 10.1063/1.4948610
    24. 24
      Larcher, D.; Palacín, M. R.; Amatucci, G. G.; Tarascon, J. M. Electrochemically Active LiCoO2 and LiNiO2 Made by Cationic Exchange Under Hydrothermal Conditions. J. Electrochem. Soc. 1997, 144, 408417,  DOI: 10.1149/1.1837424
    25. 25
      Mishra, K.; Pundir, S. S.; Rai, D. K. All-solid-state Proton Battery Using Gel Polymer Electrolyte. AIP Conf. Proc. 2014, 1591, 633,  DOI: 10.1063/1.4872700
    26. 26
      Smith, J. P.; Brown, W. E.; Lehr, J. R. Structure of Crystalline Phosphoric Acid. J. Am. Chem. Soc. 1955, 77, 27282730,  DOI: 10.1021/ja01615a013
    27. 27
      Dippel, T.; Kreuer, K.; Lassegues, J.; Rodriguez, D. Proton Conductivity in Fused Phosphoric Acid; A 1H/31P PFG-NMR and QNS Study. Solid State Ionics 1993, 61, 4146,  DOI: 10.1016/0167-2738(93)90332-W
    28. 28
      Swift, M. W.; Qi, Y. First-principles Prediction of Potentials and Space-charge Layers in All-solid-state Batteries. Phys. Rev. Lett. 2019, 122, 167701  DOI: 10.1103/PhysRevLett.122.167701
    29. 29
      Neugebauer, J.; Van de Walle, C. G. Hydrogen in GaN: Novel Aspects of a Common Impurity. Phys. Rev. Lett. 1995, 75, 44524455,  DOI: 10.1103/PhysRevLett.75.4452
    30. 30
      Fingerle, M.; Buchheit, R.; Sicolo, S.; Albe, K.; Hausbrand, R. Reaction and Space Charge Layer Formation at the LiCoO2–LiPON Interface: Insights on Defect Formation and Ion Energy Level Alignment by a Combined Surface Science–Simulation Approach. Chem. Mater. 2017, 29, 76757685,  DOI: 10.1021/acs.chemmater.7b00890
    31. 31
      Tukamoto, H.; West, A. R. Electronic Conductivity of LiCoO2 and Its Enhancement by Magnesium Doping. J. Electrochem. Soc. 1997, 144, 31643168,  DOI: 10.1149/1.1837976
    32. 32
      Momma, K.; Izumi, F. VESTA 3 for Three-dimensional Visualization of Crystal, Volumetric and Morphology Data. J. Appl. Crystallogr. 2011, 44, 12721276,  DOI: 10.1107/S0021889811038970
    33. 33
      Johnson, III, R. D. NIST101. Computational Chemistry Comparison and Benchmark Database, https://cccbdb.nist.gov/.
    34. 34
      Van de Walle, C. G.; Neugebauer, J. Universal Alignment of Hydrogen Levels in Semiconductors, Insulators and Solutions. Nature 2003, 423, 626628,  DOI: 10.1038/nature01665
    35. 35
      Xu, G.; Li, J.; Wang, C.; du, X.; Lu, D.; Xie, B.; Wang, X.; Lu, C.; Liu, H.; Dong, S.; Cui, G.; Chen, L. The Formation/Decomposition Equilibrium of LiH and Its Contribution on Anode Failure in Practical Lithium Metal Batteries. Angew. Chem., Int. Ed. 2021, 60, 77707776,  DOI: 10.1002/anie.202013812
    36. 36
      Wakabayashi, Y.; Shirasawa, T.; Voegeli, W.; Takahashi, T. Observation of Structure of Surfaces and Interfaces by Synchrotron X-ray Diffraction: Atomic-scale Imaging and Time-resolved Measurements. J. Phys. Soc. Jpn. 2018, 87, 061010  DOI: 10.7566/JPSJ.87.061010
    37. 37
      Zheng, S. J.; Fisher, C. A. J.; Hitosugi, T.; Kumatani, A.; Shiraki, S.; Ikuhara, Y. H.; Kuwabara, A.; Moriwake, H.; Oki, H.; Ikuhara, Y. Antiphase Inversion Domains in Lithium Cobaltite Thin Films Deposited on Single-crystal Sapphire Substrates. Acta Mater. 2013, 61, 76717678,  DOI: 10.1016/j.actamat.2013.09.004
    38. 38
      Blöchl, P. E. Projector Augmented-wave Method. Phys. Rev. B 1994, 50, 1795317979,  DOI: 10.1103/PhysRevB.50.17953
    39. 39
      Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865,  DOI: 10.1103/PhysRevLett.77.3865
    40. 40
      Klimeš, J.; Bowler, D. R.; Michaelides, A. Chemical Accuracy for the van der Waals Density Functional. J. Phys. Condens. Matter. 2010, 22, 022201  DOI: 10.1088/0953-8984/22/2/022201
    41. 41
      Klimeš, J.; Bowler, D. R.; Van der Michaelides, A. Waals Density Functionals Applied to Solids. Phys. Rev. B 2011, 83, 195131  DOI: 10.1103/PhysRevB.83.195131
    42. 42
      Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-energy Calculations Using a Plane-wave Basis Set. Phys. Rev. B 1996, 54, 1116911186,  DOI: 10.1103/PhysRevB.54.11169
    43. 43
      Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.; Sutton, A. P. Electron-energy-loss Spectra and the Structural Stability of Nickel Oxide: An LSDA+ U Study. Phys. Rev. B 1998, 57, 15051509,  DOI: 10.1103/PhysRevB.57.1505
    44. 44
      Freysoldt, C.; Neugebauer, J.; Van de Walle, C. G. Fully Ab Initio Finite-size Corrections for Charged-defect Supercell Calculations. Phys. Rev. Lett. 2009, 102, 016402  DOI: 10.1103/PhysRevLett.102.016402
    45. 45
      Imamoğlu, A. Cavity QED Based on Collective Magnetic Dipole Coupling: Spin Ensembles as Hybrid Two-level Systems. Phys. Rev. Lett. 2009, 102, 083602  DOI: 10.1103/PhysRevLett.102.083602
    46. 46
      Reuter, K.; Scheffler, M. Composition, Structure, and Stability of RuO2 (110) as a Function of Oxygen Pressure. Phys. Rev. B 2001, 65, 035406  DOI: 10.1103/PhysRevB.65.035406
    47. 47
      Koettgen, J.; Zacherle, T.; Grieshammer, S.; Martin, M. Ab Initio Calculation of the Attempt Frequency of Oxygen Diffusion in Pure and Samarium Doped Ceria. Phys. Chem. Chem. Phys. 2017, 19, 99579973,  DOI: 10.1039/c6cp04802a
    48. 48
      Kyrtsos, A.; Matsubara, M.; Bellotti, E. Migration Mechanisms and Diffusion Barriers of Carbon and Native Point Defects in GaN. Phys. Rev. B 2016, 93, 245201  DOI: 10.1103/PhysRevB.93.245201
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsami.1c17945.

    • Fabrication of gas-exposed thin-film batteries, X-ray crystal truncation rod (CTR) scattering analysis, estimation of the interface resistance using an equivalent circuit, electrochemical properties of the pure gas-exposed batteries, impedance spectra of the H2O-vapor-exposed battery, parameters of the equivalent circuits used for the gas-exposed batteries, capacity retention, annealing recovery of the air-exposed battery, annealing processes and performance of the H2O-vapor-exposed battery, parameters of the equivalent circuits used for the gas-exposed batteries after the annealing process, optimized geometry of a proton occupying a Li vacancy site in the LiCoO2 structure, comparison of different LCO proton interstitial sites, and the time-of-flight secondary ion mass spectroscopy spectrum for H and Li species (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.