ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES
RETURN TO ISSUEPREVResearch ArticleNEXT

Nanostructured LiMnO2 with Li3PO4 Integrated at the Atomic Scale for High-Energy Electrode Materials with Reversible Anionic Redox

  • Miho Sawamura
    Miho Sawamura
    Department of Applied Chemistry, Tokyo Denki University, 5 Senju Asahi-cho, Adachi-ku, Tokyo 120-8551, Japan
  • Sho Kobayakawa
    Sho Kobayakawa
    Department of Applied Chemistry, Tokyo Denki University, 5 Senju Asahi-cho, Adachi-ku, Tokyo 120-8551, Japan
  • Jun Kikkawa
    Jun Kikkawa
    National Institute for Materials Science (NIMS), Namiki, Tsukuba, Ibaraki 305-0044, Japan
    More by Jun Kikkawa
  • Neeraj Sharma
    Neeraj Sharma
    School of Chemistry, UNSW Sydney, Sydney, NSW 2052, Australia
  • Damian Goonetilleke
    Damian Goonetilleke
    School of Chemistry, UNSW Sydney, Sydney, NSW 2052, Australia
  • Aditya Rawal
    Aditya Rawal
    Mark Wainwright Analytical Centre, UNSW Sydney, Sydney, NSW 2052, Australia
    More by Aditya Rawal
  • Nanaka Shimada
    Nanaka Shimada
    Department of Chemistry and Life Science, Yokohama National University, 79-5 Tokiwadai, Hodogaya-ku, Yokohama, Kanagawa 240-8501, Japan
  • Kentaro Yamamoto
    Kentaro Yamamoto
    Graduate School of Human and Environmental Studies, Kyoto University, Yoshida-honmachi, Sakyo, Kyoto 606-8501, Japan
  • Rina Yamamoto
    Rina Yamamoto
    Graduate School of Human and Environmental Studies, Kyoto University, Yoshida-honmachi, Sakyo, Kyoto 606-8501, Japan
  • Yingying Zhou
    Yingying Zhou
    Graduate School of Human and Environmental Studies, Kyoto University, Yoshida-honmachi, Sakyo, Kyoto 606-8501, Japan
  • Yoshiharu Uchimoto
    Yoshiharu Uchimoto
    Graduate School of Human and Environmental Studies, Kyoto University, Yoshida-honmachi, Sakyo, Kyoto 606-8501, Japan
  • Koji Nakanishi
    Koji Nakanishi
    SR Center, Ritsumeikan University, 1-1-1 Noji-Higashi, Kusatsu, Shiga 525-8577, Japan
  • Kei Mitsuhara
    Kei Mitsuhara
    SR Center, Ritsumeikan University, 1-1-1 Noji-Higashi, Kusatsu, Shiga 525-8577, Japan
  • Koji Ohara
    Koji Ohara
    Diffraction and Scattering Division, Center for Synchrotron Radiation Research, Japan Synchrotron Radiation Research Institute (JASRI, SPring-8), Kouto, Sayo-cho, Sayo-gun, Hyogo 679-5198, Japan
    More by Koji Ohara
  • Jiwon Park
    Jiwon Park
    Department of Chemistry, Korea Advanced Institute of Science and Technology (KAIST) and KAIST Institute for NanoCentury, 291 Daehak-ro, Yuseong-gu, Daejeon 34141, Republic of Korea
    More by Jiwon Park
  • Hye Ryung Byon
    Hye Ryung Byon
    Department of Chemistry, Korea Advanced Institute of Science and Technology (KAIST) and KAIST Institute for NanoCentury, 291 Daehak-ro, Yuseong-gu, Daejeon 34141, Republic of Korea
  • Hiroaki Koga
    Hiroaki Koga
    Research Organization for Information Science and Technology (RIST), 1-18-16 Hamamatsucho, Minato-ku, Tokyo 105-0013, Japan
    Elements Strategy Initiative for Catalysts and Batteries, Kyoto University, f1-30 Goryo-Ohara, Nishikyo-ku, Kyoto 615-8245, Japan
    More by Hiroaki Koga
  • Masaki Okoshi
    Masaki Okoshi
    Research Organization for Information Science and Technology (RIST), 1-18-16 Hamamatsucho, Minato-ku, Tokyo 105-0013, Japan
    Elements Strategy Initiative for Catalysts and Batteries, Kyoto University, f1-30 Goryo-Ohara, Nishikyo-ku, Kyoto 615-8245, Japan
  • Toshiaki Ohta
    Toshiaki Ohta
    SR Center, Ritsumeikan University, 1-1-1 Noji-Higashi, Kusatsu, Shiga 525-8577, Japan
  • , and 
  • Naoaki Yabuuchi*
    Naoaki Yabuuchi
    Department of Chemistry and Life Science, Yokohama National University, 79-5 Tokiwadai, Hodogaya-ku, Yokohama, Kanagawa 240-8501, Japan
    Elements Strategy Initiative for Catalysts and Batteries, Kyoto University, f1-30 Goryo-Ohara, Nishikyo-ku, Kyoto 615-8245, Japan
    Advanced Chemical Energy Research Center, Yokohama National University, 79-5 Tokiwadai, Hodogaya-ku, Yokohama, Kanagawa 240-8501, Japan
    *E-mail: [email protected]
Cite this: ACS Cent. Sci. 2020, 6, 12, 2326–2338
Publication Date (Web):December 15, 2020
https://doi.org/10.1021/acscentsci.0c01200

Copyright © 2022 American Chemical Society. This publication is licensed under these Terms of Use.

  • Open Access

Article Views

3779

Altmetric

-

Citations

LEARN ABOUT THESE METRICS
PDF (10 MB)
Supporting Info (1)»

Abstract

Nanostructured LiMnO2 integrated with Li3PO4 was successfully synthesized by the mechanical milling route and examined as a new series of positive electrode materials for rechargeable lithium batteries. Although uniform mixing at the atomic scale between LiMnO2 and Li3PO4 was not anticipated because of the noncompatibility of crystal structures for both phases, our study reveals that phosphorus ions with excess lithium ions dissolve into nanosize crystalline LiMnO2 as first evidenced by elemental mapping using STEM-EELS combined with total X-ray scattering, solid-state NMR spectroscopy, and a theoretical ab initio study. The integrated phase features a low-crystallinity metastable phase with a unique nanostructure; the phosphorus ion located at the tetrahedral site shares faces with adjacent lithium ions at slightly distorted octahedral sites. This phase delivers a large reversible capacity of ∼320 mA h g–1 as a high-energy positive electrode material in Li cells. The large reversible capacity originated from the contribution from the anionic redox of oxygen coupled with the cationic redox of Mn ions, as evidenced by operando soft XAS spectroscopy, and the superior reversibility of the anionic redox and the suppression of oxygen loss were also found by online electrochemical mass spectroscopy. The improved reversibility of the anionic redox originates from the presence of phosphorus ions associated with the suppression of oxygen dimerization, as supported by a theoretical study. From these results, the mechanistic foundations of nanostructured high-capacity positive electrode materials were established, and further chemical and physical optimization may lead to the development of next-generation electrochemical devices.

Synopsis

Li3PO4 is successfully integrated into rocksalt LiMnO2. The integrated phase delivers a large reversible capacity with anionic redox, which is used for high-energy electrode materials of Li-ion batteries.

Introduction

ARTICLE SECTIONS
Jump To

Lithium insertion materials, which are electronic or ionic conductors, are indispensable to rechargeable lithium batteries extensively used in our daily life. The understanding of the chemistry of lithium insertion materials has rapidly developed in the past four decades, (1) and the technological progress on rechargeable lithium batteries has resulted in the electrific automobile applications. For electric vehicle applications, Ni/Co-containing layered oxides are the dominant positive electrode material, coupled with graphite as a negative electrode material. (2) Critically, the market for electric vehicles is rapidly growing worldwide, and this expansion is likely to result in the shortage of Co resources. (3) Therefore, high-energy and high-performance positive electrode materials without Co ions (and preferably without Ni ions as well) are required. In the past decade, Li2MnO3-based electrode materials have been extensively studied as potential high-capacity positive electrode materials. (4−7) Compared with other layered materials such as LiMO2, where M are transition metals, Li2MnO3-based electrodes feature an enriched lithium content that results in higher theoretical capacities (at both the electrode and battery level) because the amount of the extractable Li content is larger. Although pure Li2MnO3 is electrochemically less active (or inactive), solid-solution samples such as Li1.2Co0.13Ni0.13Mn0.54O2 deliver large reversible capacities as electrode materials. (8) Charge compensation at high levels of lithium extraction has been demonstrated by the reversible oxidation of oxide ions in addition to the conventional redox of transition metal ions. (9) However, a continuous phase transition associated with the formation of trivalent Mn ions, presumably coupled with oxygen loss on the charge, restricts its use for practical applications, (7,10) especially in such rigid layered structures.
Redox reactions of such anionic species dramatically change the requirements of the host materials and hence the material design strategies for energy storage applications. (11−13) New lithium-excess and high-capacity positive electrode materials have appeared in the past several years. (14−17) Among them, binary systems with LiMnO2 have been proposed as a new class of electrode materials, including Li3NbO4–LiMnO2, (18) Li2TiO3–LiMnO2, (7) and LiF–LiMnO2, (16,19) and these Li-excess electrode materials effectively utilize the anionic redox coupled with the cationic redox of Mn3+/Mn4+. In these cases, the end-members of Li3NbO4, Li2TiO3, and LiF are known as electrochemically inactive because of the absence of conductive electrons. Nevertheless, the crystal structures of these compounds are classified as (cation-ordered) rocksalt-type structures and are considered to feature a structural compatibility with LiMnO2, which possesses a cubic close packed (ccp) oxygen lattice. These binary systems, indeed experimentally, crystallize into cation- or anion-disordered rocksalt phases as solid solution samples of both end-members. Consequently, Li ions in these electrochemically inactive compounds are activated with anionic redox. (11,20)
In this study, Li3PO4 with a hexagonal close-packed oxygen lattice is targeted for the end member of the binary system with LiMnO2. Unlike previous studies, LiMnO2 (21) and Li3PO4 (22) feature different oxygen packings, with Li and Mn ions located at octahedral sites in LiMnO2 compared to the Li and P ions located at tetrahedral sites in Li3PO4. Therefore, the formation of a thermodynamically stable solid solution between the end members is an implausible scenario. Interestingly, a lower molar mass of Li3PO4 is advantageous in increasing the electrode theoretical capacity, and the cationic and anionic redox electrochemical potentials are expected to increase due to the inductive effect of PO43– ions. (23) Although synthesis by conventional calcination failed, presumably from the structural incompatibility of the phases, we demonstrate the formation of integrated phases of LiMnO2 and Li3PO4 that were obtained by a mechanical milling route. The milling route allows the synthesis of thermodynamically metastable phases using mechanical energy, including friction and shear stress on mixing. (24−27) The incorporation of phosphorus ions to rocksalt LiMnO2 is clearly evidenced by a combined analysis with STEM-EELS, total X-ray scattering, solid-state NMR, and theoretical ab initio studies. Moreover, anionic redox is activated in this system, leading to large reversible capacities with a higher operating voltage. From these results, the possibility of a new series of high-capacity metastable electrode materials, which are integrated at the atomic-to-nano scale, are discussed.

Results and Discussion

ARTICLE SECTIONS
Jump To

Synthesis of Materials and Detailed Phase Characterization

Figure 1a shows X-ray diffraction patterns of mixtures of LiMnO2, Li3PO4, and various compositions in (1 – x)LiMnO2·xLi3PO4 after mechanical milling. After mechanical milling, the well-defined diffraction pattern of LiMnO2 (x = 0) with a zigzag-type layered structure is completely lost, and low crystallinity phases with cation-disordered rocksalt structures are observed, consistent with our previous report. (28) In contrast, no change in the crystal structure for Li3PO4 (x = 1) was observed after milling under the same experimental conditions; however there was a noticeable decrease in the crystallinity. For the binary system, where x = 0.1 and 0.2, the diffraction pattern of Li3PO4 was completely lost, and a cation disordered rocksalt phase formed. The phase evolution process of x = 0.2 with different milling times, which was reformulated as Li7/6P1/6Mn2/3O2, is shown in Supporting Figure S1. After milling, TEM analysis revealed that a primary particle size reduced into less than 20 nm (Figure 1b and Supporting Figure S2), consistent with the observation of broad reflections in the diffraction patterns. Moreover, such nanoparticles are tightly agglomerated around each other, forming relatively larger secondary particles. Therefore, nanosized samples prepared by mechanical milling generally have a relatively small surface area. (29) Note that a uniform distribution of Mn and P is evidenced from scanning transmission electron microscopy (STEM) coupled with elemental mapping by energy dispersive X-ray spectroscopy (EDX), as shown in Supporting Figure S2. Nevertheless, when the fraction of Li3PO4 increases to x = 0.3, a reflection characteristic of Li3PO4 with low-crystallinity at 2θ = 22° appears. Phase segregation was also observed by EDX mapping, as shown in Supporting Figure S3. These results clearly indicate that the “solubility” of Li3PO4 in LiMnO2 is limited in a relatively narrow range using the milling method.

Figure 1

Figure 1. Characterization of a binary system of (1 – x)LiMnO2·xLi3PO4. (a) XRD patterns of LiMnO2 and Li3PO4 before and after mechanical milling as well as of mixtures of LiMnO2 and Li3PO4 with different compositions, (1 – x)LiMnO2·xLi3PO4, after mechanical milling. Schematic illustrations of crystal structures were drawn using the VESTA program. (59) (b) Phosphorus and manganese distributions obtained by STEM-EELS elemental mapping; a STEM image of the sample x = 0.2 in (1 – x)LiMnO2·xLi3PO4 is also shown. Observed P L-edge EELS spectra of points α and β are also shown. Averaged and full EELS spectra of the sample and the data for Li3PO4 after milling are also shown for comparison. (c) Clear lattice fringes can be seen in a high-resolution TEM image (also see Supporting Figure S7b).

Although “apparently” single-phase samples for x = 0.1 and 0.2 were successfully obtained by mechanical milling, the detailed structure of the binary phase is still not clear. To further examine the crystal structures in the LiMnO2–Li3PO4 binary system, cation distributions in the crystalline phase were determined using Rietveld analysis. Earlier work by Hayashi and co-workers on electrodes for solid-state batteries examined mechanical milling in the LiCoO2 and Li2SO4 binary system. (24) They proposed that nanosize LiCoO2 was embedded in an amorphous matrix of Li2SO4, forming a nanocomposite. Further work associated with solid-state batteries examined a sputtered thin amorphous layer of Li3PO4 on LiNiO2. (30) A question arises; amorphous Li3PO4 cannot be obtained by mechanical milling, as shown in Figure 1a, and crystalline Li3PO4 only shows a reduction in the crystallinity with milling. Notably, synchrotron XRD data show a higher crystallinity for the rocksalt phase for x = 0.2 when compared with that of x = 0 without Li3PO4 (Supporting Figure S4). Therefore, the possibility of PO43– incorporation into crystalline domains with Mn ions cannot be eliminated.
To validate this hypothesis, a detailed structure of nanosize and rocksalt LiMnO2 without phosphorus ions was first analyzed. An ideal rocksalt model in which both cations, Li and Mn ions, are located at the same octahedral 4a site is not suitable (Supporting Figure S5) as the intensity of the 111 reflection at 2θ = 12° is unmatched. The March distribution function was used along the 111 direction, resulting in a significantly better fit. Note that this model contains a preferred orientation term. For Li7/6P1/6Mn2/3O2, phosphorus ions were modeled at tetrahedral 8c sites with the preferred orientation term, and a good fitting result was obtained (Supporting Figure S5). The lattice parameter of the sample increased from a = 4.156 to 4.181 Å by phosphorus integration. This model assumes the formation of an atomic incorporation of phosphorus ions into LiMnO2 and a solid-solution-like phase, Li1+xPxMn1–2xO2. It should be noted that this is likely a metastable phase with enriched structural defects, and is therefore segregated into LiMnO2 and Li3PO4 upon heating (Supporting Figure S6a); however, the metastable phase is stable after one-year storage in the glovebox at room temperature (Supporting Figure S6b). In the solid-solution model, a large repulsive electrostatic interaction is unavoidable for face-shared sites between octahedral and tetrahedral sites (discussed later).
To clarify the structural analysis, detailed nanostructures were analyzed by STEM combined with electron energy loss spectroscopy (EELS), with atomic distributions observed with a 1.5 nm spatial resolution. Typical EELS spectra of the P L-edge are also shown in Figure 1b, and Mn L-edge and Li K-edge spectra are shown in Supporting Figures 7a; atomic distributions of P and Mn ions obtained from the intensity in the EELS spectra are also shown in Figure 1b. EELS analysis reveals that P ions are well integrated with Mn ions at the nanoscale, and uniform distributions of cationic species are also noted. The major phase appears to be a nanosize crystalline phase with a size of less than 5 nm (Figure 1b and Supporting Figure S7b), and the XRD and TEM observations and data are inconsistent with the scenario of the formation of amorphous Li3PO4 and a composite structure with crystalline LiMnO2. The redistribution of both cationic species clearly proceeds by mechanical milling in this binary system. Surface conversion reactions were also proposed for the nanocomposite of LiF and MnO, (31) but a more uniform distribution of the cationic species is noted in the binary system of LiMnO2–Li3PO4. In addition, the incorporation of P ions into the crystalline LiMnO2 phase is further supported from the profile of P L-edge EELS spectra. P L-edge EELS spectra collected from different spots are clearly different from those of Li3PO4 (Figure 1b), with a small but notable energy shift in the spectra relative to Li3PO4. A similar trend is also noted in the Li K-edge and Mn L-edge EELS spectra (Supporting Figure S7a).
The incorporation of P ions into the crystalline LiMnO2 phase is further evidenced by total X-ray scattering measurements, with patterns in reciprocal space for different samples shown in Supporting Figure S8. The reduced pair distribution functions (PDFs) for rocksalt LiMnO2 and Li7/6P1/6Mn2/3O2 are shown in Figure 2a. Although both samples show similar PDFs in the r-range over 3 Å, a clear difference is visible in first and second peaks at 1.6 and 2.0 Å, respectively. The intensity of the 2.0 Å peak originating from Mn–O bonds decreases, while that of the 1.6 Å peak increases; this originates from P–O bonds, as clearly found in reference PDF data for Li3PO4. Note that no change in the PDF data before and after milling was observed for Li3PO4 (Supporting Figure S9), indicating that amorphous Li3PO4 was not formed by mechanical milling. Li7/6P1/6Mn2/3O2 is a metastable phase (Supporting Figure S6a) that transforms to zigzag LiMnO2 and Li3PO4 during heating to 550 °C, as is clearly observed in the PDF by taking the sum of the data for LiMnO2 and Li3PO4 (Figure 2a). The clustering of PO43– ions in the 550 °C heated structure clearly influences the profile of the PDF data in the wide r-range (Figure 2a) and is in contrast with milled Li7/6P1/6Mn2/3O2 and rocksalt LiMnO2, where the wide r-range data essentially overlap. This is strong evidence that phosphorus ions are nonclustered and diluted to rocksalt LiMnO2 in the milled Li7/6P1/6Mn2/3O2. A solid solution was formed rather than clusters.

Figure 2

Figure 2. Analysis of local structures for the as-prepared sample x = 0.2, Li7/6P1/6Mn2/3O2. (a) Experimental X-ray pair distribution function (PDF), (b) Raman spectra, and (c) solid-state NMR spectra. The data for the reference materials are also shown. From the all data, the clustering of phosphorus ions is not seen for Li7/6P1/6Mn2/3O2, indicating that phosphorus is diluted to crystalline rocksalt LiMnO2.

These observations are also supported by Raman and solid-state NMR spectroscopy. Figure 2b compares the Raman spectra of LiMnO2, Li3PO4, and Li7/6P1/6Mn2/3O2. Two clearly resolved peaks were observed for the sample after milling. The peak at 623 cm–1 is assigned as the vibration of the Mn–O bonds in the MnO6 octahedra, and a similar peak is also visible for rocksalt LiMnO2 at 614 cm–1. (28) In the binary system, a minute energy shift (wavenumber) compared with that of rocksalt LiMnO2 was noted. In addition, another peak at 930 cm–1 is present, originating from the symmetric stretching mode of PO4 tetrahedra. However, a clear shift in the wavenumber was also compared with the spectrum of pure Li3PO4 (955 cm–1). Note that the peak shift to 930 cm–1 wasis observed for the high-temperature phase of Li3PO4, originating from differences in the P–O interatomic distances (1.55 Å for the low-temperature phase and 1.86 Å for the high-temperature phase). (32) The P–O distance for Li7/6P1/6Mn2/3O2 was calculated to be 1.81 Å, assuming that phosphorus ions were located at tetrahedral sites in the ccp lattice (also see Supporting Figure S12a). The P–O interatomic distance is almost the same with that of the high-temperature phase of Li3PO4; therefore, the Raman peak shift for Li7/6P1/6Mn2/3O2 is also consistent with the solid solution model. In addition, the increase of the P–O distance after mixing with LiMnO2 was also found for the PDF data (Figure 2a). Such differences in the local environments are further supported by solid-state NMR spectroscopy. 7Li and 31P NMR spectra collected, including reference materials, are shown in Supporting Figure S10. There is a positive chemical shift (36 ppm) in the 7Li NMR spectra for zigzag-type LiMnO2, originating from the interaction of electrons in Mn3+ ions via oxygen (Li–O–Mn bonding). (33) After milling, a small chemical shift for the 7Li NMR spectra was noted (Figure 2c). In the disordered rocksalt structure, a number of different local environments for Li ions are created; therefore, the total spin density transferred to Li ions from Mn ions may be canceled out. Note that a broad peak centered at 36 ppm is also present, which may originate from the presence of defects in the structure that resemble LiMnO2. No difference in the chemical shift was noted for Li3PO4 for 7Li and 31P NMR spectra before and after milling, as shown in Figure 2c and Supporting Figure S11, respectively. This is also consistent with the data obtained by Raman scattering (Supporting Figure S12b). Peak broadening was observed for Li3PO4 after milling, probably associated with the lowered crystallinity by milling. For Li7/6P1/6Mn2/3O2, the chemical shift of the 7Li NMR spectra is the same as that of rocksalt LiMnO2; however, there is a modification of chemical shift in 31P NMR spectra, indicating that there is interaction of the PO4 tetrahedra with Mn ions with d electrons. These results are also consistent with the observation from PDF analysis, EELS spectra, and Raman spectra, and it was concluded that PO43– ions are nonclustered for Li7/6P1/6Mn2/3O2 and successfully dissolved into LiMnO2.

Electrochemistry

Electrochemical properties of the binary system of LiMnO2–Li3PO4 as metastable phases were examined as positive electrode materials in Li cells. Galvanostatic charge–discharge curves of the binary system are shown in Figure 3. Rocksalt LiMnO2 shows ca. 90% of the theoretical capacity (286 mA h g–1) calculated from the Li content and Mn3+/Mn4+ redox. (28) Reversible capacities of the binary system increase relative to the parent with the addition of Li3PO4. The highest reversible capacity iwas obtained for the x = 0.1 sample of over 300 mA h g–1. The reversible capacity was slightly reduced for x = 0.2 (Li7/6P1/6Mn2/3O2). However, the reversible capacity reached nearly 300 mA h g–1, which clearly exceeds the theoretical capacity based on Mn3+/Mn4+ redox (218 mA h g–1). This fact suggests that anionic redox is activated in this system, which also supports the successful integration of LiMnO2 and Li3PO4. Li ions, which are originally found in the structure of Li3PO4, are effectively extracted by the electrochemical oxidation of oxide ions for the integrated phase (anionic redox in this system is discussed further in a later section). The initial discharge capacity is slightly larger than that of the initial charge capacity. A similar trend was also observed for nanosized samples prepared by mechanical milling, originating from the partial oxidation of the sample during milling (also see the Supporting Information). (27) Interestingly, the further enrichment of Li3PO4 in the binary system results in the reduction of the reversible capacities. This observation is also consistent with the results obtained by XRD, where phase segregation of Li3PO4 wasis observed for x = 0.3. Li ions cannot be extracted from the segregated phase because Li3PO4 is an electronic insulator. Nevertheless, the well-integrated phase, x = 0.2, shows a relatively good rate capability and delivers 190 mA h g–1 at a rate of 640 mA g–1, (Figure 3b), which is significantly better than those of rocksalt LiMnO2 prepared by mechanical milling without Li3PO4. (28) Electrochemical properties of the sample of x = 0.2 were further tested with different cutoff voltages in Li cells. Initial discharge capacities are shown in Figure 3c, and charge–discharge curves and capacity retention are also compared in Supporting Figure S13. The discharge capacity increased to ca. 320 mA h g–1 with a 5.2 V cutoff. However, cyclability as electrode materials is not acceptable, and nearly 50% of the reversible capacity was lost after 30 cycles. In contrast, good capacity retention is observed with a 4.5 V cutoff, and the sample retainsed 80% of the reversible capacity. Although the reversible capacity wasis reduced to 180 mA h g–1, excellent capacity retention was achieved with a 4.1 V cutoff. Note that the reversible capacities and capacity retention are also significantly influenced by the operating temperature (Supporting Figure S14). A large reversible capacity of 320 mA h g–1 was observed at 50 °C, but a deterioration of the capacity retention was observed. In contrast, good capacity retention wasis realized at 0 °C even though the reversible capacity was reduced to 235 mA h g–1. This observation suggests that irreversible oxygen loss occurs on electrochemical cycles and is accelerated at elevated temperatures, as observed in Li2MnO3-based electrode materials. (7)

Figure 3

Figure 3. Electrochemical properties of the (1 – x)LiMnO2·xLi3PO4 binary system. (a) Galvanostatic charge–discharge curves of the samples with different compositions prepared by mechanical milling, (b) rate-capability of the sample x = 0.2, Li7/6P1/6Mn2/3O2, and (c) dependency of cutoff voltages on eversible capacities and capacity retention. The electrode loading was ca. 6.0–6.5 mg cm–2 for these electrochemical tests.

Charge Compensation Mechanisms

Reaction mechanisms of the binary system of LiMnO2–Li3PO4 as electrode materials were examined by synchrotron X-ray diffraction (SXRD) combined with hard–soft X-ray absorption spectroscopy (XAS) and online gas analysis focusing on x = 0.2, Li7/6–yP1/6Mn2/3O2. On charge, a clear shift of the reflection positions and a contraction of the crystal lattice were noted as shown in Figure 4a, indicating that a topotactic reaction proceeds. All the Bragg diffraction lines continuously shift to higher angles on charge to 220 mA h g–1 (Li extraction), and the lattice parameters change from 4.178 to 4.061 Å. However, only a minor change in the reflection positions and the lattice parameter were noted for the further charge to 4.8 V (i.e., similar lattice parameters were noted for the fully charged sample and the sample charged to 220 mA h g–1). On discharge, reflection broadening was clearly observed, and the lattice parameters of the fully discharged sample were much larger than those of the as-prepared sample, suggesting that irreversible reactions occur on electrochemical cycles (at least for the first cycle). The origin of said degradation is further discussed in a later section.

Figure 4

Figure 4. Reaction mechanisms of Li7/6–yP1/6Mn2/3O2. (a) Changes in synchrotron XRD patterns. (b and c) Charge compensation processes observed by an operando soft XAS study shown by (b) a schematic illustration of the experimental setup, (34) and (c) variations in the Mn L-edge and O K-edge XAS spectra. (d and e) In situ partial gas pressures of CO2, H2, and O2 and the corresponding charge curves from nanosized Li2MnO3 and Li7/6P1/6Mn2/3O2, respectively. The dashed line is the baseline of the gas pressure, and the filled area indicates the change of the partial gas pressure through the gas evolution.

To understand the charge compensation process, soft X-ray absorption spectroscopy was applied to study changes in electronic structures based on previous work indicating oxygen redox activity for Li-excess electrode materials using this technique. (15) However, ex-situ soft XAS spectroscopy requires high-vacuum conditions that, when coupled with possible extrinsic changes induced in the samples during washing, drying, and long exposure to vacuum, can provide misleading information. Therefore, operando soft X-ray absorption spectroscopy was developed (34,35) and utilized in this study (Figure 4b). Instead of a conventional carbonate electrolyte, lithium bis(trifluoromethanesulfonyl)amide (TFSA) dissolved in acetonitrile was utilized as the electrolyte to reduce X-ray absorption. Although TFSA anions contain oxygen in the structure, the major absorption was observed above 533.5 eV, as shown in Supporting Figure S15a; therefore, changes in O K-edge XAS spectra near pre-edge regions from the electrode materials could be clearly detected. Aluminum-sputtered Si3N4 was used as windows to allow the transmittance of the synchrotron X-rays. Acetonitrile is reactive with metallic Li; therefore, a delithiated phase of lithium iron phosphate, Li0.5FePO4, was used as a counter electrode. A charge curve taken during the operando measurement is shown in Supporting Figure S15b. The irreversible capacity was relatively large, probably originating from electrolyte decomposition. Therefore, the chemical composition and amount of extracted Li ions (equaling the charge capacity) were estimated and calibrated by comparing the cutoff voltages. A clear shift of the O K-edge spectra was noted on charge, as shown in Figure 4c, and a similar trend was observed with the ex situ study shown in Supporting Figure S16. After charging to 110 mA h g–1 (corresponding to 4 V vs Li+/Li), Mn was in the tetravalent state, and no further change was observed beyond the 4 V region for the oxidation state of Mn ions. Pronounced changes in the O K-edge XAS spectra were subsequently noted, and peak intensity around 532 eV increased as the charge capacity and voltage increased. A similar trend is noted with other Li-excess materials with Mn ions. (7) This peak is proposed to originate from the hole creation and stabilization by a π-type interaction with Mn t2g orbital. (36) The appearance of a clear peak at 532 eV is noted for Li1.2–yTi0.4Mn0.4O2, (15) suggesting that the stability of the holes created by electrochemical oxidation is energetically less stable for the nanosize samples prepared by mechanical milling. This finding is also consistent with the non-negligible deterioration of cyclability after charging to 5.2 V and at elevated temperatures (Supporting Figures S13 and S14). Significantly better capacity retention was achieved with a 4.5 V cutoff as shown in Figure 3c. The changes in electronic structures of phosphorus ions were also studied, and P K-edge XAS spectra were collected, as shown in Supporting Figure S16d. A gradual shift of the spectra to a higher energy region was noted on the charge process. A similar phenomenon was also observed for Li1–xFePO4, (37) and the peak shift was expected to originate from changes in local structures (shrinkage of the crystal lattice) and bonding characters associated with the oxidation of Mn ions. These changes in P K-edge spectra also support the successful integration of phosphorus ions into LiMnO2.
Note that on the discharge process the partial Mn reduction to a divalent state, associated with minor oxygen loss on charge, is evidenced by the ex situ XAS study (Supporting Figure S16). Such phenomena are generally observed in Li2MnO3-based high-capacity electrode materials. (7) Nevertheless, the profiles of O K-edge XAS spectra for second charge–discharge are superimposed with those for first charge–discharge as shown in Supporting Figure S17, indicating the high reversibility in the anionic redox for Li7/6P1/6Mn2/3O2 on the charge–discharge processes.
The suppression of oxygen loss for the phosphorus-integrated composition is also supported by online electrochemical mass spectrometry (OEMS) in Figure 4d and e. (38) For comparison, nanosized Li2MnO3 has been prepared by mechanical milling (Supporting Figure S18). For nanosized Li2MnO3, the CO2 partial pressure drastically increases on lithium extraction (Figure 4d), which is consistent with published data. (39) CO2 is considered to originate from the side reaction between the electrolyte solvent and the highly active singlet oxygen that is released from the electrode materials by irreversible anionic redox. (40) Irreversible oxygen loss from the active material is a dominative process for nanosized Li2MnO3. In addition, hydrogen gas was detected as the charge capacity and voltage increased, which was previously detected for Li1.2–yMn0.4Ti0.4O2 on charge and likely originated from the oxidative decomposition of the electrolyte solution at >4.7 V. (35) In contrast, gas evolution for both CO2 and H2 was notably suppressed for Li7/6–yP1/6Mn2/3O2 on charge to 4.5 V (Figure 4e), although a small amount of CO2 was detected in the higher voltage region; this amount was significantly smaller than nanosized Li2MnO3. This finding is consistent with the better electrochemical reversibility for Li7/6P1/6Mn2/3O2 with a 4.5 V cutoff (Figure 3c). To further increase the cyclability as electrode materials with anionic redox, a surface coating, (41) functional binder, (42) electrolyte additive, (43) and high-voltage-tolerated electrolyte (44) would be effective to suppress unfavorable side reactions with electrolyte.

Theoretical Understanding of Crystal Structure and Reaction Mechanisms

To provide a unified picture of the phosphorus integration into LiMnO2, theoretical calculations were conducted. Models used were derived from the experimental data that P ions are diluted in LiMnO2 and no P-based clusters occur. First, a large unit cell of Li16Mn16O32 was constructed, and then two Mn ions were replaced by Li and P ions, leading to the formation of Li17Mn14PO32. After structural relaxation, the formation of the unique nanostructure was noted. The P ion at the tetrahedral site shares faces with three lithium ions originally located at octahedral sites. An electrostatic repulsive interaction between P and Li ions results in the displacement of Li ions from the original octahedral position, as shown in Figure 5a. The total energy of Li17Mn14PO32 is 3.11 eV higher than that of a simple mixture of Li16Mn16O32 (16 units of LiMnO2) and Li3PO4, which is consistent with the fact that the P-integrated phase is the metastable phase (Supporting Figure S6). Nevertheless, the total energy is relatively stabilized through the structural relaxation associated with the displacement of the face-shared Li ions. Note that when the trivalent Mn ion is forced to share the face with the P ion instead of the Li ion, the energy is further destabilized by 1.4 eV. These results indicate that the formation of the Li-rich phase and the short-range ordering of one P and three Li ions are necessary to energetically stabilize such unique metastable phases.

Figure 5

Figure 5. Theoretical calculation of Li7/6P1/6Mn2/3O2. (a) An optimized crystal structure of rocksalt LiMnO2 after P-integration, and a local structure of the P ion in the rocksalt phase. Changes in the number of valence electrons for (b) Mn, P, and (c) O ions upon the delithiation process from Li9Mn6PO16 (∼Li7/6P1/6Mn2/3O2). Li ions are extracted from the model one by one, which is the lowest energy in the structure. Electron density changes during delithiation are also visualized in panel d. Yellow and cyan isosurfaces (drawn at 0.01 e Bohr–1) (3) indicate electron accumulation and depletion, respectively; □ denotes vacant sites. The data for −3, – 7, and −9 Li are also shown in Supporting Figure S20.

Li extraction processes from the P-integrated LiMnO2 were further studied by theoretical calculations, and the model with a smaller unit cell of Li9Mn6PO16 was used (also see the Supporting Information). Note that the unit cell volume of Li9Mn6PO16 is larger (approximately 2.3%) than that of Li8Mn8O16 without phosphorus ions, and a similar trend in the unit cell volume change was observed from experimental observations (Supporting Figure S4). The calculated voltage upon Li extraction (Supporting Figure S19) shows two different regionsat 3 and 4 V, which likely correspond to cationic Mn and anionic O redox, respectively. Changes in the oxidation states of the P-integrated phase were evaluated by the Bader method; the number of valence electrons on each atomic site upon Li extraction summarized is Figure 5b and c, and variations in the charge density are visualized in Figure 5d and Supporting Figure S20. From the valence electrons, oxide ions are classified as oxygen bonded or oxygen nonbonded to phosphorus. The covalency of the P–O bonds is much higher than those of the Li–O and Mn–O bonds, resulting in the difference in the number of valence electrons. On the charge process, Mn ions are oxidized and coupled with Li extraction; therefore, the number of valence electrons is reduced for Mn ions, forming □6Li3Mn6PO16 (□ denotes vacant sites). Upon further oxidation and Li extraction, a clear reduction of the number of valence electrons was observed for oxygen (Figure 5c). This finding is consistent with experimentally observed operando XAS data (Figure 4c). Note that oxygen bonded to phosphorus is not responsible for charge compensation and oxygen bonded to Mn ions is selectively oxidized, as shown in the number of valence electron and differential charge density distributions. Moreover, no dimerization of oxygen for the fully charged sample, □9Mn6PO16, was observed even after structural relaxation. It is proposed that the presence of phosphorus ions in the structure suppresses the “excess” oxidation of oxide ions; therefore, the possibility of oxygen dimerization around phosphorus ions would be reduced in this system, leading to the better reversibility of anionic redox. This is consistent with better performance for the P-integrated phase relative to that of the nanosized Li2MnO3 demonstrated above.

Conclusions

ARTICLE SECTIONS
Jump To

In this study, a new series of electrode materials has been proposed, i.e., rocksalt-type metastable crystalline phases with atomically incorporated P ions. The successful integration of phosphorus ions with excess Li ions into nanosize and crystalline LiMnO2 was demonstrated by the combination of different characterization techniques. The P ion is located at the tetrahedral site in the CCP lattice, which shares faces with adjacent Li ions, and a slight distortion from the ideal octahedral site for Li sites energetically stabilizes this unique nanostructure. This phase delivers a large reversible capacity of ∼320 mA h g–1 that originates from the contribution from cationic redox reaction of Mn ions coupled with oxygen anionic redox, and the superior reversibility of the anionic redox is realized by the presence of the P ion in this phase. The use of such materials with unique nanostructures is a conceptually new approach to electrode design, which may lead to the next generation of materials. This metastable system opens the possibility for a new path to design high-capacity electrode materials. Moreover, this new concept for material design could also extended to different systems, such as sodium or potassium battery applications, leading to the development of a sustainable energy society in the future.

Experimental Section

ARTICLE SECTIONS
Jump To

Synthesis of Materials

A nanoscale integrated LiMnO2–Li3PO4 binary system was prepared by mechanical milling using a planetary ball mill (PULVERISETTE 7; FRITSCH). For the sample of x = 0.2 in the (1 – x)LiMnO2·xLi3PO4 binary system, 1.146 g of LiMnO2 and 0.354 g of Li3PO4 (97%, Wako) were initially mixed using a zirconia pot (45 mL) and zirconia balls (14.8 g) at 600 rpm for 12 h. After milling for 12 h, the mixture was removed from the container and mixed with a mortar and pestle to ensure sample uniformity during the milling. The mixture was again milled using the zirconia pot and balls at 600 rpm for 12 h. This process was repeated three times, and the mixture was milled for 36 h in total. LiMnO2 was prepared from Li2CO3 (98.5%, Kanto Kagaku) and Mn2O3 at 900 °C for 12 h in an Ar atmosphere. Mn2O3 was obtained by heating MnCO3 (Kishida Chemical Co., Ltd.) at 850 °C for 12 h in air. After the preparation, the samples were stored in an Ar-filled glovebox. The sample handling before and after the mechanical milling was also carried out in an Ar-filled glovebox without exposure to moist air.

Electrochemical Measurement

After synthetic milling, the samples were further mixed with acetylene black (HS-100; Denka Co., Ltd., LiMnO2–Li3PO4/AB 90:10 wt %) using the planetary ball mill at 300 rpm for 12 h with the zirconia container and balls. Better electrode performance was achieved through this process even though a relatively higher content of carbon sacrifices the volumetric energy density. Composite positive electrodes comprised of 76.5 wt % LiMnO2–Li3PO4, 13.5 wt % acetylene black, and 10 wt % polyvinylidene fluoride (KF 1100; Kureha Co. Ltd.) dispersed in N-methylpyrrolidone were pasted on aluminum foil as a current collector. The electrodes were dried at 80 °C for 2 h in vacuum and then heated at 120 °C for 2 h. Metallic lithium (Honjo Metal Co., Ltd.) was used as a negative electrode. The electrolyte solution used was 1.0 mol dm–3 LiPF6 dissolved in ethylene carbonate/dimethyl carbonate (3:7 by volume, battery grade; Kishida Chemical Corp., Ltd.). Two-electrode cells (TJ-AC; Tomcell Japan) were assembled in an Ar-filled glovebox. The cells were cycled at a rate of 10 mA g–1 at room temperature.

Characterization of Materials

X-ray diffraction (XRD) patterns of the samples were collected using an X-ray diffractometer (D2 PHASER; Bruker Corp., Ltd.) equipped with a one-dimensional X-ray detector using Cu Ka radiation generated at 300 W (30 kV and 10 mA) with a Ni filter. An airtight sample holder was used for the XRD measurement. Synchrotron XRD data were collected on the beamline BL16XU at SPring-8 synchrotron facility in Japan. The measurement was conducted using an automatic powder diffraction system for the Debye–Scherrer geometry by placing the sample in a glass capillary. The wavelength used was 0.5003 Å, which was calibrated with CeO2 as a reference sample. Structural analysis was carried out using RIETAN-FP. (45)
STEM–EELS was conducted using an electron microscope (Titan Cubed; Thermo Fisher Scientific) equipped with a spectroscope (GIF Quantum ERS; Gatan). The electron beam energy and EELS resolution were 300 keV and 1.3 eV, respectively. The P/Mn and Li/Mn atomic ratios were calculated with P-L2,3, Mn-M2,3, and Li–K edges at each pixel using software (DigitalMicrograph; Gatan). Calculated Li/Mn ratios were corrected using the similarly calculated ratios of the reference materials, Li2MnO3 and LiMnO2.
X-ray total scattering measurements was performed with an incident X-ray energy of E = 61.4 keV at the BL04B2 beamline in SPring-8, Japan, to study the local structure by pair distribution function (PDF) analysis. Hybrid detectors of Ge and CdTe were employed for the data collection of the total structure factor S(Q). The reduced PDF G(r) was obtained by the conventional Fourier transform of S(Q). (46)
The solid-state NMR spectra were acquired on a Bruker Avance III 300 MHz spectrometer with a 7 T superconducting magnet operating at a frequency of 116.6 MHz for the 7Li nucleus and 121.4 MHz for the 31P nucleus. Pristine powders of the positive electrode materials were packed into 2.5 mm zirconia rotors in an argon-filled glovebox before being transferred to the spectrometer. The rotors were spun to 30 kHz MAS under a nitrogen atmosphere. Single-pulse excitation was used to detect the 7Li NMR spectra with a 2 μs hard pulse for excitation and 1 s recycle delays. The spectra were referenced to the 7Li spectrum of solid LiF set to 0 ppm. For 31P NMR, the spectra were acquired with a 2.65 μs 90° pulse and a 120 s recycle delay. The 31P chemical shift was referenced using the 31P signal of NH4H2PO4 set to 1 ppm.
Raman spectra of the samples were collected by using a Raman spectrometer (NRS-4100, JASCO Ltd.) with a 532 nm laser. The measurement was conducted with a sample holder (LIBcell; Nanophoton) without exposure to air.
Hard X-ray absorption spectroscopy (XAS) at the Mn K-edge was performed at beamline BL-12C of the Photon Factory Synchrotron Source in Japan. Hard XAS spectra were collected with a silicon monochromator in the transmission mode. The intensities of the incident and transmitted X-rays were measured using an ionization chamber at room temperature. Composite electrode samples were prepared using the two-electrode cells at a rate of 10 mA g–1. The composite electrodes were rinsed with dimethyl carbonate and sealed in a water-resistant polymer film in the Ar-filled glovebox. Normalization of the XAS spectra was carried out using the program code IFEFFIT. (47) The postedge background was determined using a cubic spline procedure.
The ex situ and operando XAS measurements for O K-edge and Mn L-edge were collected at BL27SU in SPring-8, Japan. For ex situ measurements, the same sample preparation method was used as that for the hard XAS measurement. The absorption spectra were collected with the partial fluorescence mode. The detailed experimental setup of the operando XAS measurement is found in literature. (34)
An online mass electrochemical mass spectrometry (OEMS) system was established as described in previous work. (38) The custom-designed OEMS cell (Tomcell Japan) was comprised of electrode materials, a lithium foil as the negative electrode, and 1 M LiPF6 in a mixture of ethylene carbonate (EC) and dimethyl carbonate (DMC, 3:7 by volume) as the electrolyte solution. The assembled OEMS cell was rested in an Ar-filled glovebox for 2 h and linked to the OEMS apparatus in a constant-temperature container maintained at 25 °C. The OEMS cell was purged with an Ar gas (99.9999%) for a total 10 min to test the system. The cell pressure (16–18 psi) and the mass spectrum were also monitored at an open-circuit voltage at room temperature. The CO2 gas evolution from DMC evaporation stabilized after 3 h, which was then set as background. The galvanostatic charging process was carried out at a current density of 10 mA g–1. The gas analysis was conducted after collecting gas products every 1 h.
Spin-polarized DFT + U calculations were performed using the GPU-port (48,49) of the Vienna Ab Initio Simulation Package (50,51) (VASP). Exchange–correlation effects and ion–electron interactions were described by the Perdew–Burke–Ernzerhof (PBE) functional (52) and the projector-augmented wave (PAW) method, (53,54) respectively. Dispersion interactions were described by DFT-D3. (55) The cutoff energy for plane-wave expansion was 520 eV, and an onsite Coulomb interaction correction of UJ = 3.9 eV (56) was applied to Mn 3d using the method by Dudarev et al. (57) In structural relaxation calculations, atomic coordinates as well as cell parameters were optimized. The Bader method was used to determine the number of valence electrons on each atomic site. (58)

Safety Statement

No unexpected or unusually high-safety hazards were encountered.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acscentsci.0c01200.

  • DFT calculation details; structural evolutions of the LiMnO2–Li3PO4 system; STEM images; EDX maps; SEM images; comparison of SXRD patterns; Reitveld analysis of rocksalt LiMnO2 and Li7/6P1/6Mn2/3O2; phase segregation of Li7/6P1/6Mn2/3O2 after heating; analysis of the Mn distrivution; X-ray diffraction patterns, X-ray pair distribution functions; Raman, XAS, and solid-state NMR spectra; charge–discharge and charge curves; charge density differences, and synthesis and electrochemical properties of nanosize rocksalt Li2MnO3 (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

ARTICLE SECTIONS
Jump To

  • Corresponding Author
    • Naoaki Yabuuchi - Department of Chemistry and Life Science, Yokohama National University, 79-5 Tokiwadai, Hodogaya-ku, Yokohama, Kanagawa 240-8501, JapanElements Strategy Initiative for Catalysts and Batteries, Kyoto University, f1-30 Goryo-Ohara, Nishikyo-ku, Kyoto 615-8245, JapanAdvanced Chemical Energy Research Center, Yokohama National University, 79-5 Tokiwadai, Hodogaya-ku, Yokohama, Kanagawa 240-8501, JapanOrcidhttp://orcid.org/0000-0002-9404-5693 Email: [email protected]
  • Authors
    • Miho Sawamura - Department of Applied Chemistry, Tokyo Denki University, 5 Senju Asahi-cho, Adachi-ku, Tokyo 120-8551, Japan
    • Sho Kobayakawa - Department of Applied Chemistry, Tokyo Denki University, 5 Senju Asahi-cho, Adachi-ku, Tokyo 120-8551, Japan
    • Jun Kikkawa - National Institute for Materials Science (NIMS), Namiki, Tsukuba, Ibaraki 305-0044, Japan
    • Neeraj Sharma - School of Chemistry, UNSW Sydney, Sydney, NSW 2052, AustraliaOrcidhttp://orcid.org/0000-0003-1197-6343
    • Damian Goonetilleke - School of Chemistry, UNSW Sydney, Sydney, NSW 2052, AustraliaPresent Address: D.G. Battery and Electrochemistry Laboratory (BELLA), Institute of Nanotechnology, Karlsruhe Institute for Technology, Hermann-von-Helmholtz-Platz 1, 76344 Eggenstein-Leopoldshafen, GermanyOrcidhttp://orcid.org/0000-0003-1033-4787
    • Aditya Rawal - Mark Wainwright Analytical Centre, UNSW Sydney, Sydney, NSW 2052, AustraliaOrcidhttp://orcid.org/0000-0002-5396-1265
    • Nanaka Shimada - Department of Chemistry and Life Science, Yokohama National University, 79-5 Tokiwadai, Hodogaya-ku, Yokohama, Kanagawa 240-8501, Japan
    • Kentaro Yamamoto - Graduate School of Human and Environmental Studies, Kyoto University, Yoshida-honmachi, Sakyo, Kyoto 606-8501, JapanOrcidhttp://orcid.org/0000-0002-8739-4246
    • Rina Yamamoto - Graduate School of Human and Environmental Studies, Kyoto University, Yoshida-honmachi, Sakyo, Kyoto 606-8501, Japan
    • Yingying Zhou - Graduate School of Human and Environmental Studies, Kyoto University, Yoshida-honmachi, Sakyo, Kyoto 606-8501, Japan
    • Yoshiharu Uchimoto - Graduate School of Human and Environmental Studies, Kyoto University, Yoshida-honmachi, Sakyo, Kyoto 606-8501, JapanOrcidhttp://orcid.org/0000-0002-1491-2647
    • Koji Nakanishi - SR Center, Ritsumeikan University, 1-1-1 Noji-Higashi, Kusatsu, Shiga 525-8577, Japan
    • Kei Mitsuhara - SR Center, Ritsumeikan University, 1-1-1 Noji-Higashi, Kusatsu, Shiga 525-8577, Japan
    • Koji Ohara - Diffraction and Scattering Division, Center for Synchrotron Radiation Research, Japan Synchrotron Radiation Research Institute (JASRI, SPring-8), Kouto, Sayo-cho, Sayo-gun, Hyogo 679-5198, JapanOrcidhttp://orcid.org/0000-0002-3134-512X
    • Jiwon Park - Department of Chemistry, Korea Advanced Institute of Science and Technology (KAIST) and KAIST Institute for NanoCentury, 291 Daehak-ro, Yuseong-gu, Daejeon 34141, Republic of KoreaOrcidhttp://orcid.org/0000-0002-2258-310X
    • Hye Ryung Byon - Department of Chemistry, Korea Advanced Institute of Science and Technology (KAIST) and KAIST Institute for NanoCentury, 291 Daehak-ro, Yuseong-gu, Daejeon 34141, Republic of KoreaOrcidhttp://orcid.org/0000-0003-3692-6713
    • Hiroaki Koga - Research Organization for Information Science and Technology (RIST), 1-18-16 Hamamatsucho, Minato-ku, Tokyo 105-0013, JapanElements Strategy Initiative for Catalysts and Batteries, Kyoto University, f1-30 Goryo-Ohara, Nishikyo-ku, Kyoto 615-8245, Japan
    • Masaki Okoshi - Research Organization for Information Science and Technology (RIST), 1-18-16 Hamamatsucho, Minato-ku, Tokyo 105-0013, JapanElements Strategy Initiative for Catalysts and Batteries, Kyoto University, f1-30 Goryo-Ohara, Nishikyo-ku, Kyoto 615-8245, Japan
    • Toshiaki Ohta - SR Center, Ritsumeikan University, 1-1-1 Noji-Higashi, Kusatsu, Shiga 525-8577, Japan
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

N.Y. acknowledges the partial support from JSPS, Grant-in-Aid for Scientific Research (Grants 15H05701, 18H02076, and 19H05816), and the MEXT program “Elements Strategy Initiative to Form Core Research Center (no. JPMXP0112101003)” from MEXT, Ministry of Education Culture, Sports, Science and Technology, Japan. N.Sharma and D.G. thanks the Australian Research Council (ARC) for support through the projects DE160100237, DP170100269, and the research training scheme. K.O. thanks the support from JSPS, Grant-in-Aid for Scientific Research (Grant 19H05814). H.R.B. acknowledges partial support from the KAIST-funded Global Singularity Research Program. The synchrotron X-ray absorption work was done under the approval of the Photon Factory Program Advisory Committee (Proposal no. 2019G033). The synchrotron radiation experiments were performed at the BL04B2, BL19B2, and BL27SU of SPring-8 with the approval of the Japan Synchrotron Radiation Research Institute (JASRI) (Proposal nos. 2018B1811, 2019A2058, 2016A1460, and 2016A1463). The TEM observation was supported by Dr. Keisuke Shinoda, NIMS, Battery Research Platform with support by NIMS microstructural characterization platform as a program of the “Nanotechnology Platform” of the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan. The computational resource of the AI Bridging Cloud Infrastructure (ABCI) provided by National Institute of Advanced Industrial Science and Technology (AIST) was used. N.Y. also thanks Prof. Masanori Tachikawa, Assoc. Prof. Takayoshi Ishimoto, and Assist. Prof. Makito Takagi (Yokohama City University, Japan) for the fruitful discussion about structural models for the theoretical calculation.

References

ARTICLE SECTIONS
Jump To

This article references 59 other publications.

  1. 1
    Whittingham, M. S. Ultimate Limits to Intercalation Reactions for Lithium Batteries. Chem. Rev. 2014, 114, 1141411443,  DOI: 10.1021/cr5003003
  2. 2
    Lee, W.; Muhammad, S.; Kim, T.; Kim, H.; Lee, E.; Jeong, M.; Son, S.; Ryou, J.-H.; Yoon, W.-S. New Insight into Ni-Rich Layered Structure for Next-Generation Li Rechargeable Batteries. Adv. Energy Mater. 2018, 8, 1701788,  DOI: 10.1002/aenm.201701788
  3. 3
    Olivetti, E. A.; Ceder, G.; Gaustad, G. G.; Fu, X. Lithium-Ion Battery Supply Chain Considerations: Analysis of Potential Bottlenecks in Critical Metals. Joule 2017, 1, 229243,  DOI: 10.1016/j.joule.2017.08.019
  4. 4
    Johnson, C. S.; Kim, J. S.; Lefief, C.; Li, N.; Vaughey, J. T.; Thackeray, M. M. The significance of the Li2MnO3 component in ’composite’ xLi2MnO3 center dot (1-x)LiMn0.5Ni0.5O2 electrodes. Electrochem. Commun. 2004, 6, 10851091,  DOI: 10.1016/j.elecom.2004.08.002
  5. 5
    Lu, Z. H.; MacNeil, D. D.; Dahn, J. R. Layered cathode materials Li[NixLi(1/3–2x/3)Mn(2/3-x/3)]O2 for lithium-ion batteries. Electrochem. Solid-State Lett. 2001, 4, A191A194,  DOI: 10.1149/1.1407994
  6. 6
    Robertson, A. D.; Bruce, P. G. Mechanism of electrochemical activity in Li2MnO3. Chem. Mater. 2003, 15, 19841992,  DOI: 10.1021/cm030047u
  7. 7
    Yabuuchi, N.; Kubota, K.; Aoki, Y.; Komaba, S. Understanding Particle-Size-Dependent Electrochemical Properties of Li2MnO3-Based Positive Electrode Materials for Rechargeable Lithium Batteries. J. Phys. Chem. C 2016, 120, 875885,  DOI: 10.1021/acs.jpcc.5b10517
  8. 8
    Johnson, C. S.; Li, N.; Lefief, C.; Vaughey, J. T.; Thackeray, M. M. Synthesis, Characterization and Electrochemistry of Lithium Battery Electrodes: xLi2MnO3·(1 – x)LiMn0.333Ni0.333Co0.333O2 (0 ≤ x ≤ 0.7). Chem. Mater. 2008, 20, 60956106,  DOI: 10.1021/cm801245r
  9. 9
    Koga, H.; Croguennec, L.; Ménétrier, M.; Douhil, K.; Belin, S.; Bourgeois, L.; Suard, E.; Weill, F.; Delmas, C. Reversible Oxygen Participation to the Redox Processes Revealed for Li1.20Mn0.54Co0.13Ni0.13O2. J. Electrochem. Soc. 2013, 160, A786A792,  DOI: 10.1149/2.038306jes
  10. 10
    Chen, C.-J.; Pang, W. K.; Mori, T.; Peterson, V. K.; Sharma, N.; Lee, P.-H.; Wu, S.-h.; Wang, C.-C.; Song, Y.-F.; Liu, R.-S. The Origin of Capacity Fade in the Li2MnO3·LiMO2 (M = Li, Ni, Co, Mn) Microsphere Positive Electrode: An Operando Neutron Diffraction and Transmission X-ray Microscopy Study. J. Am. Chem. Soc. 2016, 138, 88248833,  DOI: 10.1021/jacs.6b03932
  11. 11
    Yabuuchi, N. Material Design Concept of Lithium-Excess Electrode Materials with Rocksalt-Related Structures for Rechargeable Non-Aqueous Batteries. Chem. Rec. 2019, 19 (4), 690,  DOI: 10.1002/tcr.201800089
  12. 12
    Ben Yahia, M.; Vergnet, J.; Saubanère, M.; Doublet, M.-L. Unified picture of anionic redox in Li/Na-ion batteries. Nat. Mater. 2019, 18, 496502,  DOI: 10.1038/s41563-019-0318-3
  13. 13
    Sudayama, T.; Uehara, K.; Mukai, T.; Asakura, D.; Shi, X.-M.; Tsuchimoto, A.; Mortemard de Boisse, B.; Shimada, T.; Watanabe, E.; Harada, Y.; Nakayama, M.; Okubo, M.; Yamada, A. Multiorbital bond formation for stable oxygen-redox reaction in battery electrodes. Energy Environ. Sci. 2020, 13 (5), 1492,  DOI: 10.1039/C9EE04197D
  14. 14
    Zhao, W.; Yamaguchi, K.; Sato, T.; Yabuuchi, N. Li4/3Ni1/3Mo1/3O2 – LiNi1/2Mn1/2O2 Binary System as High Capacity Positive Electrode Materials for Rechargeable Lithium Batteries. J. Electrochem. Soc. 2018, 165, A1357A1362,  DOI: 10.1149/2.0661807jes
  15. 15
    Yabuuchi, N.; Nakayama, M.; Takeuchi, M.; Komaba, S.; Hashimoto, Y.; Mukai, T.; Shiiba, H.; Sato, K.; Kobayashi, Y.; Nakao, A.; Yonemura, M.; Yamanaka, K.; Mitsuhara, K.; Ohta, T. Origin of Stabilization and Destabilization in Solid-State Redox Reaction of Oxide Ions for Rechargeable Lithium Batteries. Nat. Commun. 2016, 7 (1), 13814,  DOI: 10.1038/ncomms13814
  16. 16
    House, R. A.; Jin, L.; Maitra, U.; Tsuruta, K.; Somerville, J. W.; Förstermann, D. P.; Massel, F.; Duda, L.; Roberts, M. R.; Bruce, P. G. Lithium manganese oxyfluoride as a new cathode material exhibiting oxygen redox. Energy Environ. Sci. 2018, 11, 926932,  DOI: 10.1039/C7EE03195E
  17. 17
    Lee, J.; Kitchaev, D. A.; Kwon, D.-H.; Lee, C.-W.; Papp, J. K.; Liu, Y.-S.; Lun, Z.; Clément, R. J.; Shi, T.; McCloskey, B. D.; Guo, J.; Balasubramanian, M.; Ceder, G. Reversible Mn2+/Mn4+ double redox in lithium-excess cathode materials. Nature 2018, 556, 185190,  DOI: 10.1038/s41586-018-0015-4
  18. 18
    Yabuuchi, N.; Takeuchi, M.; Nakayama, M.; Shiiba, H.; Ogawa, M.; Nakayama, K.; Ohta, T.; Endo, D.; Ozaki, T.; Inamasu, T.; Sato, K.; Komaba, S. High-capacity electrode materials for rechargeable lithium batteries: Li3NbO4-based system with cation-disordered rocksalt structure. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 76507655,  DOI: 10.1073/pnas.1504901112
  19. 19
    Ji, H.; Wu, J.; Cai, Z.; Liu, J.; Kwon, D.-H.; Kim, H.; Urban, A.; Papp, J. K.; Foley, E.; Tian, Y.; Balasubramanian, M.; Kim, H.; Clément, R. J.; McCloskey, B. D.; Yang, W.; Ceder, G. Ultrahigh power and energy density in partially ordered lithium-ion cathode materials. Nat. Energy 2020, 5, 213221,  DOI: 10.1038/s41560-020-0573-1
  20. 20
    Kobayashi, Y.; Sawamura, M.; Kondo, S.; Harada, M.; Noda, Y.; Nakayama, M.; Kobayakawa, S.; Zhao, W.; Nakao, A.; Yasui, A.; Rajendra, H. B.; Yamanaka, K.; Ohta, T.; Yabuuchi, N. Activation and stabilization mechanisms of anionic redox for Li storage applications: Joint experimental and theoretical study on Li2TiO3–LiMnO2 binary system. Mater. Today 2020, 37, 43,  DOI: 10.1016/j.mattod.2020.03.002
  21. 21
    Croguennec, L.; Deniard, P.; Brec, R. Electrochemical Cyclability of Orthorhombic LiMnO2: Characterization of Cycled Materials. J. Electrochem. Soc. 1997, 144, 33233330,  DOI: 10.1149/1.1838013
  22. 22
    Keffer, C.; Mighell, A.; Mauer, F.; Swanson, H.; Block, S. CRYSTAL STRUCTURE OF TWINNED LOW-TEMPERATURE LITHIUM PHOSPHATE. Inorg. Chem. 1967, 6, 119,  DOI: 10.1021/ic50047a027
  23. 23
    Padhi, A. K.; Nanjundaswamy, K. S.; Masquelier, C.; Okada, S.; Goodenough, J. B. Effect of Structure on the Fe3 + /Fe2 + Redox Couple in Iron Phosphates. J. Electrochem. Soc. 1997, 144, 16091613,  DOI: 10.1149/1.1837649
  24. 24
    Nagao, K.; Hayashi, A.; Deguchi, M.; Tsukasaki, H.; Mori, S.; Tatsumisago, M. Amorphous LiCoO2Li2SO4 active materials: Potential positive electrodes for bulk-type all-oxide solid-state lithium batteries with high energy density. J. Power Sources 2017, 348, 18,  DOI: 10.1016/j.jpowsour.2017.02.038
  25. 25
    Kitajou, A.; Kobayashi, E.; Okada, S. Electrochemical Performance of a Novel Cathode material “LiFeOF” for Li-ion Batteries. Electrochemistry 2015, 83, 885888,  DOI: 10.5796/electrochemistry.83.885
  26. 26
    Chen, R. Y.; Ren, S. H.; Knapp, M.; Wang, D.; Witter, R.; Fichtner, M.; Hahn, H. Disordered Lithium-Rich Oxyfluoride as a Stable Host for Enhanced Li+ Intercalation Storage. Adv. Energy Mater. 2015, 5, 1401814,  DOI: 10.1002/aenm.201401814
  27. 27
    Hoshino, S.; Glushenkov, A. M.; Ichikawa, S.; Ozaki, T.; Inamasu, T.; Yabuuchi, N. Reversible Three-Electron Redox Reaction of Mo3+/Mo6+ for Rechargeable Lithium Batteries. ACS Energy Letters 2017, 2, 733738,  DOI: 10.1021/acsenergylett.7b00037
  28. 28
    Sato, T.; Sato, K.; Zhao, W.; Kajiya, Y.; Yabuuchi, N. Metastable and nanosize cation-disordered rocksalt-type oxides: revisit of stoichiometric LiMnO2 and NaMnO2. J. Mater. Chem. A 2018, 6, 1394313951,  DOI: 10.1039/C8TA03667E
  29. 29
    Tsuchiya, Y.; Glushenkov, A. M.; Yabuuchi, N. Effect of Nanosizing on Reversible Sodium Storage in a NaCrO2 Electrode. ACS Applied Nano Materials 2018, 1, 364370,  DOI: 10.1021/acsanm.7b00207
  30. 30
    Sabi, Y.; Sato, S.; Hayashi, S.; Furuya, T.; Kusanagi, S. A new class of amorphous cathode active material LixMyPOz (M = Ni, Cu, Co, Mn, Au, Ag, Pd). J. Power Sources 2014, 258, 5460,  DOI: 10.1016/j.jpowsour.2014.02.021
  31. 31
    Jung, S.-K.; Kim, H.; Cho, M. G.; Cho, S.-P.; Lee, B.; Kim, H.; Park, Y.-U.; Hong, J.; Park, K.-Y.; Yoon, G.; Seong, W. M.; Cho, Y.; Oh, M. H.; Kim, H.; Gwon, H.; Hwang, I.; Hyeon, T.; Yoon, W.-S.; Kang, K. Lithium-free transition metal monoxides for positive electrodes in lithium-ion batteries. Nat. Energy 2017, 2, 16208,  DOI: 10.1038/nenergy.2016.208
  32. 32
    Popović, L.; Manoun, B.; de Waal, D.; Nieuwoudt, M. K.; Comins, J. D. Raman spectroscopic study of phase transitions in Li3PO4. J. Raman Spectrosc. 2003, 34, 7783,  DOI: 10.1002/jrs.954
  33. 33
    Lee, Y. J.; Grey, C. P. 6Li Magic-Angle Spinning (MAS) NMR Study of Electron Correlations, Magnetic Ordering, and Stability of Lithium Manganese(III) Oxides. Chem. Mater. 2000, 12, 38713878,  DOI: 10.1021/cm000469t
  34. 34
    Hattori, M.; Yamamoto, K.; Matsui, M.; Nakanishi, K.; Mandai, T.; Choudhary, A.; Tateyama, Y.; Sodeyama, K.; Uchiyama, T.; Orikasa, Y.; Tamenori, Y.; Takeguchi, T.; Kanamura, K.; Uchimoto, Y. Role of Coordination Structure of Magnesium Ions on Charge and Discharge Behavior of Magnesium Alloy Electrode. J. Phys. Chem. C 2018, 122, 2520425210,  DOI: 10.1021/acs.jpcc.8b08558
  35. 35
    Yamamoto, K.; Zhou, Y.; Yabuuchi, N.; Nakanishi, K.; Yoshinari, T.; Kobayashi, T.; Kobayashi, Y.; Yamamoto, R.; Watanabe, A.; Orikasa, Y.; Tsuruta, K.; Park, J.; Byon, H. R.; Tamenori, Y.; Ohta, T.; Uchimoto, Y. Charge Compensation Mechanism of Lithium-Excess Metal Oxides with Different Covalent and Ionic Characters Revealed by Operando Soft and Hard X-ray Absorption Spectroscopy. Chem. Mater. 2020, 32, 139147,  DOI: 10.1021/acs.chemmater.9b02838
  36. 36
    Okubo, M.; Yamada, A. Molecular Orbital Principles of Oxygen-Redox Battery Electrodes. ACS Appl. Mater. Interfaces 2017, 9, 3646336472,  DOI: 10.1021/acsami.7b09835
  37. 37
    Yoon, W.-S.; Chung, K. Y.; McBreen, J.; Zaghib, K.; Yang, X.-Q. Electronic Structure of the Electrochemically Delithiated Li[sub 1–x]FePO[sub 4] Electrodes Investigated by P K-edge X-Ray Absorption Spectroscopy. Electrochem. Solid-State Lett. 2006, 9, A415,  DOI: 10.1149/1.2216619
  38. 38
    Bonnet-Mercier, N.; Wong, R. A.; Thomas, M. L.; Dutta, A.; Yamanaka, K.; Yogi, C.; Ohta, T.; Byon, H. R. A structured three-dimensional polymer electrolyte with enlarged active reaction zone for Li–O2 batteries. Sci. Rep. 2015, 4, 7127,  DOI: 10.1038/srep07127
  39. 39
    Rana, J.; Papp, J. K.; Lebens-Higgins, Z.; Zuba, M.; Kaufman, L. A.; Goel, A.; Schmuch, R.; Winter, M.; Whittingham, M. S.; Yang, W.; McCloskey, B. D.; Piper, L. F. J. Quantifying the Capacity Contributions during Activation of Li2MnO3. ACS Energy Letters 2020, 5, 634641,  DOI: 10.1021/acsenergylett.9b02799
  40. 40
    Wandt, J.; Freiberg, A. T. S.; Ogrodnik, A.; Gasteiger, H. A. Singlet oxygen evolution from layered transition metal oxide cathode materials and its implications for lithium-ion batteries. Mater. Today 2018, 21, 825833,  DOI: 10.1016/j.mattod.2018.03.037
  41. 41
    Hu, S.; Li, Y.; Chen, Y.; Peng, J.; Zhou, T.; Pang, W. K.; Didier, C.; Peterson, V. K.; Wang, H.; Li, Q.; Guo, Z. Insight of a Phase Compatible Surface Coating for Long-Durable Li-Rich Layered Oxide Cathode. Adv. Energy Mater. 2019, 9, 1901795,  DOI: 10.1002/aenm.201901795
  42. 42
    Zhang, S.; Gu, H.; Pan, H.; Yang, S.; Du, W.; Li, X.; Gao, M.; Liu, Y.; Zhu, M.; Ouyang, L.; Jian, D.; Pan, F. A Novel Strategy to Suppress Capacity and Voltage Fading of Li- and Mn-Rich Layered Oxide Cathode Material for Lithium-Ion Batteries. Adv. Energy Mater. 2017, 7, 1601066,  DOI: 10.1002/aenm.201601066
  43. 43
    Hekmatfar, M.; Kazzazi, A.; Eshetu, G. G.; Hasa, I.; Passerini, S. Understanding the Electrode/Electrolyte Interface Layer on the Li-Rich Nickel Manganese Cobalt Layered Oxide Cathode by XPS. ACS Appl. Mater. Interfaces 2019, 11, 4316643179,  DOI: 10.1021/acsami.9b14389
  44. 44
    Doi, T.; Shimizu, Y.; Matsumoto, R.; Hashinokuchi, M.; Inaba, M. Suppression of Mn–Ion-Dissolution of LiNi0.5Mn1.5O4 Electrodes in a Highly Concentrated Electrolyte Solution at Elevated Temperatures. ChemistrySelect 2017, 2, 88248827,  DOI: 10.1002/slct.201701668
  45. 45
    Izumi, F.; Momma, K. Three-Dimensional Visualization in Powder Diffraction. Solid State Phenom. 2007, 130, 1520,  DOI: 10.4028/www.scientific.net/SSP.130.15
  46. 46
    Ohara, K.; Tominaka, S.; Yamada, H.; Takahashi, M.; Yamaguchi, H.; Utsuno, F.; Umeki, T.; Yao, A.; Nakada, K.; Takemoto, M.; Hiroi, S.; Tsuji, N.; Wakihara, T. Time-resolved pair distribution function analysis of disordered materials on beamlines BL04B2 and BL08W at SPring-8. J. Synchrotron Radiat. 2018, 25, 16271633,  DOI: 10.1107/S1600577518011232
  47. 47
    Newville, M. IFEFFIT: interactive XAFS analysis and FEFF fitting. J. Synchrotron Radiat. 2001, 8, 322324,  DOI: 10.1107/S0909049500016964
  48. 48
    Hacene, M.; Anciaux-Sedrakian, A.; Rozanska, X.; Klahr, D.; Guignon, T.; Fleurat-Lessard, P. Accelerating VASP electronic structure calculations using graphic processing units. J. Comput. Chem. 2012, 33, 25812589,  DOI: 10.1002/jcc.23096
  49. 49
    Hutchinson, M.; Widom, M. VASP on a GPU: Application to exact-exchange calculations of the stability of elemental boron. Comput. Phys. Commun. 2012, 183, 14221426,  DOI: 10.1016/j.cpc.2012.02.017
  50. 50
    Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47, 558561,  DOI: 10.1103/PhysRevB.47.558
  51. 51
    Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 1116911186,  DOI: 10.1103/PhysRevB.54.11169
  52. 52
    Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
  53. 53
    Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 50, 1795317979,  DOI: 10.1103/PhysRevB.50.17953
  54. 54
    Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 59, 17581775,  DOI: 10.1103/PhysRevB.59.1758
  55. 55
    Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104,  DOI: 10.1063/1.3382344
  56. 56
    Jain, A.; Hautier, G.; Ong, S. P.; Moore, C. J.; Fischer, C. C.; Persson, K. A.; Ceder, G. Formation enthalpies by mixing GGA and GGA + U calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 84, 045115  DOI: 10.1103/PhysRevB.84.045115
  57. 57
    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: Condens. Matter Mater. Phys. 1998, 57, 15051509,  DOI: 10.1103/PhysRevB.57.1505
  58. 58
    Henkelman, G.; Arnaldsson, A.; Jónsson, H. A fast and robust algorithm for Bader decomposition of charge density. Comput. Mater. Sci. 2006, 36, 354360,  DOI: 10.1016/j.commatsci.2005.04.010
  59. 59
    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

Cited By

This article is cited by 20 publications.

  1. Fumihiko Ichihara, Kodai Niitsu, Yoshinori Tanaka, Yasuhiro Niwa, Kazutaka Mitsuishi, Shogo Miyoshi, Takahisa Ohno, Takuya Masuda. Structural Analysis of the LiCoPO4 Electrode/NASICON-Type Li1.3Al0.3Ti1.7(PO4)3 Solid Electrolyte Interface. The Journal of Physical Chemistry C 2023, 127 (31) , 15043-15050. https://doi.org/10.1021/acs.jpcc.3c02132
  2. Asuka Kanno, Yosuke Ugata, Issei Ikeuchi, Mitsuhiro Hibino, Kensuke Nakura, Yuka Miyaoka, Izuru Kawamura, Daisuke Shibata, Toshiaki Ohta, Naoaki Yabuuchi. Durable Manganese-Based Li-Excess Electrode Material without Voltage Decay: Metastable and Nanosized Li2MnO1.5F1.5. ACS Energy Letters 2023, 8 (6) , 2753-2761. https://doi.org/10.1021/acsenergylett.3c00372
  3. Yanjia Zhang, Miho Sawamura, Maho Harada, Yusuke Noda, Masanobu Nakayama, Masato Goto, Daisuke Kan, Yuichi Shimakawa, Benoît Denis Louis Campéon, Daisuke Shibata, Toshiaki Ohta, Naoaki Yabuuchi. Partially Reversible Anionic Redox for Lithium-Excess Cobalt Oxides with Cation-Disordered Rocksalt Structure. The Journal of Physical Chemistry C 2023, 127 (5) , 2194-2203. https://doi.org/10.1021/acs.jpcc.2c08005
  4. Ryutaro Fukuma, Maho Harada, Wenwen Zhao, Miho Sawamura, Yusuke Noda, Masanobu Nakayama, Masato Goto, Daisuke Kan, Yuichi Shimakawa, Masao Yonemura, Naohiro Ikeda, Ryuta Watanuki, Henrik L. Andersen, Anita M. D’Angelo, Neeraj Sharma, Jiwon Park, Hye Ryung Byon, Sayuri Fukuyama, Zhenji Han, Hitoshi Fukumitsu, Martin Schulz-Dobrick, Keisuke Yamanaka, Hirona Yamagishi, Toshiaki Ohta, Naoaki Yabuuchi. Unexpectedly Large Contribution of Oxygen to Charge Compensation Triggered by Structural Disordering: Detailed Experimental and Theoretical Study on a Li3NbO4–NiO Binary System. ACS Central Science 2022, 8 (6) , 775-794. https://doi.org/10.1021/acscentsci.2c00238
  5. Yuji Mahara, Naoyuki Nagasako, Hideaki Oka, Yasuhito Kondo, Satoru Kosaka, Hiroyuki Nakano, Takamasa Nonaka, Yoshinari Makimura. How Fluorine Introduction Solves the Spinel Transition, a Fundamental Problem of Mn-Based Positive Electrodes. ACS Applied Materials & Interfaces 2022, 14 (21) , 24321-24331. https://doi.org/10.1021/acsami.2c02868
  6. Zhenpeng Yao, Maria K. Y. Chan, Chris Wolverton. Exploring the Origin of Anionic Redox Activity in Super Li-Rich Iron Oxide-Based High-Energy-Density Cathode Materials. Chemistry of Materials 2022, 34 (10) , 4536-4547. https://doi.org/10.1021/acs.chemmater.2c00322
  7. Miyuki Shinoda, Hongahally Basappa Rajendra, Naoaki Yabuuchi. Rocksalt and Layered Metal Sulfides for Li Storage Applications: LiMe0.5Ti0.5S2 (Me = Fe2+, Mn2+, and Mg2+). ACS Applied Energy Materials 2022, 5 (3) , 2642-2646. https://doi.org/10.1021/acsaem.1c04044
  8. Ruijie Qi, Itsuki Konuma, Benoît D. L. Campéon, Yuko Kaneda, Masashi Kondo, Naoaki Yabuuchi. Highly Graphitic Carbon Coating on Li1.25Nb0.25V0.5O2 Derived from a Precursor with a Perylene Core for High-Power Battery Applications. Chemistry of Materials 2022, 34 (4) , 1946-1955. https://doi.org/10.1021/acs.chemmater.1c04426
  9. Yuxuan Ji, Jian Wei, Di Liang, Bing Chen, Xueting Li, Hao Zhang, Zongyou Yin. LiCoO 2 cathode surface modification with optimally structured Li 3 PO 4 for outstanding high-voltage cycling performance. Nanoscale 2023, 15 (28) , 11898-11908. https://doi.org/10.1039/D3NR01251D
  10. Nanaka Shimada, Yosuke Ugata, Satoshi Nishikawa, Daisuke Shibata, Toshiaki Ohta, Naoaki Yabuuchi. Improved electrode reversibility of anionic redox with highly concentrated electrolyte solution and aramid-coated polyolefin separator. Energy Advances 2023, 2 (4) , 508-512. https://doi.org/10.1039/D3YA00066D
  11. Yanjia ZHANG, Benoît D. L. CAMPÉON, Naoaki YABUUCHI. Improved Electrode Reversibility of Nanosized Li1.15Nb0.15Mn0.7O2 through Li3PO4 Integration. Electrochemistry 2023, 91 (3) , 037004-037004. https://doi.org/10.5796/electrochemistry.23-00011
  12. Hideshi Uematsu, Nozomu Ishiguro, Masaki Abe, Shuntaro Takazawa, Jungmin Kang, Itsuki Konuma, Naoaki Yabuuchi, Yukio Takahashi. Nanoscale domain imaging of Li-rich disordered rocksalt-type cathode materials with X-ray spectroscopic ptychography. Physical Chemistry Chemical Physics 2023, 25 (5) , 3867-3874. https://doi.org/10.1039/D2CP04087E
  13. Maria Diaz-Lopez, Philip A. Chater, Olivier Proux, Yves Joly, Jean-Louis Hazemann, Pierre Bordet, Valerie Pralong. Li trapping in nanolayers of cation ‘disordered’ rock salt cathodes. Journal of Materials Chemistry A 2022, 10 (34) , 17415-17423. https://doi.org/10.1039/D2TA04262B
  14. Naoaki Yabuuchi. Rational material design of Li-excess metal oxides with disordered rock salt structure. Current Opinion in Electrochemistry 2022, 34 , 100978. https://doi.org/10.1016/j.coelec.2022.100978
  15. Neung Kwon, Akira Nasu, Atsushi Sakuda, Masahiro Tatsumisago, Akitoshi Hayashi. Amorphous positive electrode materials prepared using LiMn 1.5 Ni 0.5 O 4 and lithium oxyacid salts. Chemistry Letters 2022, https://doi.org/10.1246/cl.220132
  16. Ruijie QI, Benoît D. L. CAMPÉON, Itsuki KONUMA, Yoshihiko SATO, Yuko KANEDA, Masashi KONDO, Naoaki YABUUCHI. Metastable and Nanosized Li1.2Nb0.2V0.6O2 for High-Energy Li-ion Batteries. Electrochemistry 2022, 90 (3) , 037005-037005. https://doi.org/10.5796/electrochemistry.22-00005
  17. Benoît Denis Louis Campéon, Naoaki Yabuuchi. Fundamentals of metal oxide/oxyfluoride electrodes for Li-/Na-ion batteries. Chemical Physics Reviews 2021, 2 (4) , 041306. https://doi.org/10.1063/5.0052741
  18. Chen Xu, Siqi Guan, Lixiang Li, Chengguo Sun, Baigang An, Xin Geng. Electrochemical Properties of LiNi0.6Co0.2Mn0.2O2 Cathode Materials Prepared with Different Ammonia Content. Coatings 2021, 11 (8) , 932. https://doi.org/10.3390/coatings11080932
  19. Naohiro Ikeda, Itsuki Konuma, Hongahally Basappa Rajendra, Taira Aida, Naoaki Yabuuchi. Why is the O3 to O1 phase transition hindered in LiNiO 2 on full delithiation?. Journal of Materials Chemistry A 2021, 9 (29) , 15963-15967. https://doi.org/10.1039/D1TA03066C
  20. Takuro Sato, Kazuki Yoshikawa, Wenwen Zhao, Tokio Kobayashi, Hongahally Basappa Rajendra, Masao Yonemura, Naoaki Yabuuchi. Efficient Stabilization of Na Storage Reversibility by Ti Integration into O′3-Type NaMnO 2. Energy Material Advances 2021, 2021 https://doi.org/10.34133/2021/9857563
  • Abstract

    Figure 1

    Figure 1. Characterization of a binary system of (1 – x)LiMnO2·xLi3PO4. (a) XRD patterns of LiMnO2 and Li3PO4 before and after mechanical milling as well as of mixtures of LiMnO2 and Li3PO4 with different compositions, (1 – x)LiMnO2·xLi3PO4, after mechanical milling. Schematic illustrations of crystal structures were drawn using the VESTA program. (59) (b) Phosphorus and manganese distributions obtained by STEM-EELS elemental mapping; a STEM image of the sample x = 0.2 in (1 – x)LiMnO2·xLi3PO4 is also shown. Observed P L-edge EELS spectra of points α and β are also shown. Averaged and full EELS spectra of the sample and the data for Li3PO4 after milling are also shown for comparison. (c) Clear lattice fringes can be seen in a high-resolution TEM image (also see Supporting Figure S7b).

    Figure 2

    Figure 2. Analysis of local structures for the as-prepared sample x = 0.2, Li7/6P1/6Mn2/3O2. (a) Experimental X-ray pair distribution function (PDF), (b) Raman spectra, and (c) solid-state NMR spectra. The data for the reference materials are also shown. From the all data, the clustering of phosphorus ions is not seen for Li7/6P1/6Mn2/3O2, indicating that phosphorus is diluted to crystalline rocksalt LiMnO2.

    Figure 3

    Figure 3. Electrochemical properties of the (1 – x)LiMnO2·xLi3PO4 binary system. (a) Galvanostatic charge–discharge curves of the samples with different compositions prepared by mechanical milling, (b) rate-capability of the sample x = 0.2, Li7/6P1/6Mn2/3O2, and (c) dependency of cutoff voltages on eversible capacities and capacity retention. The electrode loading was ca. 6.0–6.5 mg cm–2 for these electrochemical tests.

    Figure 4

    Figure 4. Reaction mechanisms of Li7/6–yP1/6Mn2/3O2. (a) Changes in synchrotron XRD patterns. (b and c) Charge compensation processes observed by an operando soft XAS study shown by (b) a schematic illustration of the experimental setup, (34) and (c) variations in the Mn L-edge and O K-edge XAS spectra. (d and e) In situ partial gas pressures of CO2, H2, and O2 and the corresponding charge curves from nanosized Li2MnO3 and Li7/6P1/6Mn2/3O2, respectively. The dashed line is the baseline of the gas pressure, and the filled area indicates the change of the partial gas pressure through the gas evolution.

    Figure 5

    Figure 5. Theoretical calculation of Li7/6P1/6Mn2/3O2. (a) An optimized crystal structure of rocksalt LiMnO2 after P-integration, and a local structure of the P ion in the rocksalt phase. Changes in the number of valence electrons for (b) Mn, P, and (c) O ions upon the delithiation process from Li9Mn6PO16 (∼Li7/6P1/6Mn2/3O2). Li ions are extracted from the model one by one, which is the lowest energy in the structure. Electron density changes during delithiation are also visualized in panel d. Yellow and cyan isosurfaces (drawn at 0.01 e Bohr–1) (3) indicate electron accumulation and depletion, respectively; □ denotes vacant sites. The data for −3, – 7, and −9 Li are also shown in Supporting Figure S20.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 59 other publications.

    1. 1
      Whittingham, M. S. Ultimate Limits to Intercalation Reactions for Lithium Batteries. Chem. Rev. 2014, 114, 1141411443,  DOI: 10.1021/cr5003003
    2. 2
      Lee, W.; Muhammad, S.; Kim, T.; Kim, H.; Lee, E.; Jeong, M.; Son, S.; Ryou, J.-H.; Yoon, W.-S. New Insight into Ni-Rich Layered Structure for Next-Generation Li Rechargeable Batteries. Adv. Energy Mater. 2018, 8, 1701788,  DOI: 10.1002/aenm.201701788
    3. 3
      Olivetti, E. A.; Ceder, G.; Gaustad, G. G.; Fu, X. Lithium-Ion Battery Supply Chain Considerations: Analysis of Potential Bottlenecks in Critical Metals. Joule 2017, 1, 229243,  DOI: 10.1016/j.joule.2017.08.019
    4. 4
      Johnson, C. S.; Kim, J. S.; Lefief, C.; Li, N.; Vaughey, J. T.; Thackeray, M. M. The significance of the Li2MnO3 component in ’composite’ xLi2MnO3 center dot (1-x)LiMn0.5Ni0.5O2 electrodes. Electrochem. Commun. 2004, 6, 10851091,  DOI: 10.1016/j.elecom.2004.08.002
    5. 5
      Lu, Z. H.; MacNeil, D. D.; Dahn, J. R. Layered cathode materials Li[NixLi(1/3–2x/3)Mn(2/3-x/3)]O2 for lithium-ion batteries. Electrochem. Solid-State Lett. 2001, 4, A191A194,  DOI: 10.1149/1.1407994
    6. 6
      Robertson, A. D.; Bruce, P. G. Mechanism of electrochemical activity in Li2MnO3. Chem. Mater. 2003, 15, 19841992,  DOI: 10.1021/cm030047u
    7. 7
      Yabuuchi, N.; Kubota, K.; Aoki, Y.; Komaba, S. Understanding Particle-Size-Dependent Electrochemical Properties of Li2MnO3-Based Positive Electrode Materials for Rechargeable Lithium Batteries. J. Phys. Chem. C 2016, 120, 875885,  DOI: 10.1021/acs.jpcc.5b10517
    8. 8
      Johnson, C. S.; Li, N.; Lefief, C.; Vaughey, J. T.; Thackeray, M. M. Synthesis, Characterization and Electrochemistry of Lithium Battery Electrodes: xLi2MnO3·(1 – x)LiMn0.333Ni0.333Co0.333O2 (0 ≤ x ≤ 0.7). Chem. Mater. 2008, 20, 60956106,  DOI: 10.1021/cm801245r
    9. 9
      Koga, H.; Croguennec, L.; Ménétrier, M.; Douhil, K.; Belin, S.; Bourgeois, L.; Suard, E.; Weill, F.; Delmas, C. Reversible Oxygen Participation to the Redox Processes Revealed for Li1.20Mn0.54Co0.13Ni0.13O2. J. Electrochem. Soc. 2013, 160, A786A792,  DOI: 10.1149/2.038306jes
    10. 10
      Chen, C.-J.; Pang, W. K.; Mori, T.; Peterson, V. K.; Sharma, N.; Lee, P.-H.; Wu, S.-h.; Wang, C.-C.; Song, Y.-F.; Liu, R.-S. The Origin of Capacity Fade in the Li2MnO3·LiMO2 (M = Li, Ni, Co, Mn) Microsphere Positive Electrode: An Operando Neutron Diffraction and Transmission X-ray Microscopy Study. J. Am. Chem. Soc. 2016, 138, 88248833,  DOI: 10.1021/jacs.6b03932
    11. 11
      Yabuuchi, N. Material Design Concept of Lithium-Excess Electrode Materials with Rocksalt-Related Structures for Rechargeable Non-Aqueous Batteries. Chem. Rec. 2019, 19 (4), 690,  DOI: 10.1002/tcr.201800089
    12. 12
      Ben Yahia, M.; Vergnet, J.; Saubanère, M.; Doublet, M.-L. Unified picture of anionic redox in Li/Na-ion batteries. Nat. Mater. 2019, 18, 496502,  DOI: 10.1038/s41563-019-0318-3
    13. 13
      Sudayama, T.; Uehara, K.; Mukai, T.; Asakura, D.; Shi, X.-M.; Tsuchimoto, A.; Mortemard de Boisse, B.; Shimada, T.; Watanabe, E.; Harada, Y.; Nakayama, M.; Okubo, M.; Yamada, A. Multiorbital bond formation for stable oxygen-redox reaction in battery electrodes. Energy Environ. Sci. 2020, 13 (5), 1492,  DOI: 10.1039/C9EE04197D
    14. 14
      Zhao, W.; Yamaguchi, K.; Sato, T.; Yabuuchi, N. Li4/3Ni1/3Mo1/3O2 – LiNi1/2Mn1/2O2 Binary System as High Capacity Positive Electrode Materials for Rechargeable Lithium Batteries. J. Electrochem. Soc. 2018, 165, A1357A1362,  DOI: 10.1149/2.0661807jes
    15. 15
      Yabuuchi, N.; Nakayama, M.; Takeuchi, M.; Komaba, S.; Hashimoto, Y.; Mukai, T.; Shiiba, H.; Sato, K.; Kobayashi, Y.; Nakao, A.; Yonemura, M.; Yamanaka, K.; Mitsuhara, K.; Ohta, T. Origin of Stabilization and Destabilization in Solid-State Redox Reaction of Oxide Ions for Rechargeable Lithium Batteries. Nat. Commun. 2016, 7 (1), 13814,  DOI: 10.1038/ncomms13814
    16. 16
      House, R. A.; Jin, L.; Maitra, U.; Tsuruta, K.; Somerville, J. W.; Förstermann, D. P.; Massel, F.; Duda, L.; Roberts, M. R.; Bruce, P. G. Lithium manganese oxyfluoride as a new cathode material exhibiting oxygen redox. Energy Environ. Sci. 2018, 11, 926932,  DOI: 10.1039/C7EE03195E
    17. 17
      Lee, J.; Kitchaev, D. A.; Kwon, D.-H.; Lee, C.-W.; Papp, J. K.; Liu, Y.-S.; Lun, Z.; Clément, R. J.; Shi, T.; McCloskey, B. D.; Guo, J.; Balasubramanian, M.; Ceder, G. Reversible Mn2+/Mn4+ double redox in lithium-excess cathode materials. Nature 2018, 556, 185190,  DOI: 10.1038/s41586-018-0015-4
    18. 18
      Yabuuchi, N.; Takeuchi, M.; Nakayama, M.; Shiiba, H.; Ogawa, M.; Nakayama, K.; Ohta, T.; Endo, D.; Ozaki, T.; Inamasu, T.; Sato, K.; Komaba, S. High-capacity electrode materials for rechargeable lithium batteries: Li3NbO4-based system with cation-disordered rocksalt structure. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 76507655,  DOI: 10.1073/pnas.1504901112
    19. 19
      Ji, H.; Wu, J.; Cai, Z.; Liu, J.; Kwon, D.-H.; Kim, H.; Urban, A.; Papp, J. K.; Foley, E.; Tian, Y.; Balasubramanian, M.; Kim, H.; Clément, R. J.; McCloskey, B. D.; Yang, W.; Ceder, G. Ultrahigh power and energy density in partially ordered lithium-ion cathode materials. Nat. Energy 2020, 5, 213221,  DOI: 10.1038/s41560-020-0573-1
    20. 20
      Kobayashi, Y.; Sawamura, M.; Kondo, S.; Harada, M.; Noda, Y.; Nakayama, M.; Kobayakawa, S.; Zhao, W.; Nakao, A.; Yasui, A.; Rajendra, H. B.; Yamanaka, K.; Ohta, T.; Yabuuchi, N. Activation and stabilization mechanisms of anionic redox for Li storage applications: Joint experimental and theoretical study on Li2TiO3–LiMnO2 binary system. Mater. Today 2020, 37, 43,  DOI: 10.1016/j.mattod.2020.03.002
    21. 21
      Croguennec, L.; Deniard, P.; Brec, R. Electrochemical Cyclability of Orthorhombic LiMnO2: Characterization of Cycled Materials. J. Electrochem. Soc. 1997, 144, 33233330,  DOI: 10.1149/1.1838013
    22. 22
      Keffer, C.; Mighell, A.; Mauer, F.; Swanson, H.; Block, S. CRYSTAL STRUCTURE OF TWINNED LOW-TEMPERATURE LITHIUM PHOSPHATE. Inorg. Chem. 1967, 6, 119,  DOI: 10.1021/ic50047a027
    23. 23
      Padhi, A. K.; Nanjundaswamy, K. S.; Masquelier, C.; Okada, S.; Goodenough, J. B. Effect of Structure on the Fe3 + /Fe2 + Redox Couple in Iron Phosphates. J. Electrochem. Soc. 1997, 144, 16091613,  DOI: 10.1149/1.1837649
    24. 24
      Nagao, K.; Hayashi, A.; Deguchi, M.; Tsukasaki, H.; Mori, S.; Tatsumisago, M. Amorphous LiCoO2Li2SO4 active materials: Potential positive electrodes for bulk-type all-oxide solid-state lithium batteries with high energy density. J. Power Sources 2017, 348, 18,  DOI: 10.1016/j.jpowsour.2017.02.038
    25. 25
      Kitajou, A.; Kobayashi, E.; Okada, S. Electrochemical Performance of a Novel Cathode material “LiFeOF” for Li-ion Batteries. Electrochemistry 2015, 83, 885888,  DOI: 10.5796/electrochemistry.83.885
    26. 26
      Chen, R. Y.; Ren, S. H.; Knapp, M.; Wang, D.; Witter, R.; Fichtner, M.; Hahn, H. Disordered Lithium-Rich Oxyfluoride as a Stable Host for Enhanced Li+ Intercalation Storage. Adv. Energy Mater. 2015, 5, 1401814,  DOI: 10.1002/aenm.201401814
    27. 27
      Hoshino, S.; Glushenkov, A. M.; Ichikawa, S.; Ozaki, T.; Inamasu, T.; Yabuuchi, N. Reversible Three-Electron Redox Reaction of Mo3+/Mo6+ for Rechargeable Lithium Batteries. ACS Energy Letters 2017, 2, 733738,  DOI: 10.1021/acsenergylett.7b00037
    28. 28
      Sato, T.; Sato, K.; Zhao, W.; Kajiya, Y.; Yabuuchi, N. Metastable and nanosize cation-disordered rocksalt-type oxides: revisit of stoichiometric LiMnO2 and NaMnO2. J. Mater. Chem. A 2018, 6, 1394313951,  DOI: 10.1039/C8TA03667E
    29. 29
      Tsuchiya, Y.; Glushenkov, A. M.; Yabuuchi, N. Effect of Nanosizing on Reversible Sodium Storage in a NaCrO2 Electrode. ACS Applied Nano Materials 2018, 1, 364370,  DOI: 10.1021/acsanm.7b00207
    30. 30
      Sabi, Y.; Sato, S.; Hayashi, S.; Furuya, T.; Kusanagi, S. A new class of amorphous cathode active material LixMyPOz (M = Ni, Cu, Co, Mn, Au, Ag, Pd). J. Power Sources 2014, 258, 5460,  DOI: 10.1016/j.jpowsour.2014.02.021
    31. 31
      Jung, S.-K.; Kim, H.; Cho, M. G.; Cho, S.-P.; Lee, B.; Kim, H.; Park, Y.-U.; Hong, J.; Park, K.-Y.; Yoon, G.; Seong, W. M.; Cho, Y.; Oh, M. H.; Kim, H.; Gwon, H.; Hwang, I.; Hyeon, T.; Yoon, W.-S.; Kang, K. Lithium-free transition metal monoxides for positive electrodes in lithium-ion batteries. Nat. Energy 2017, 2, 16208,  DOI: 10.1038/nenergy.2016.208
    32. 32
      Popović, L.; Manoun, B.; de Waal, D.; Nieuwoudt, M. K.; Comins, J. D. Raman spectroscopic study of phase transitions in Li3PO4. J. Raman Spectrosc. 2003, 34, 7783,  DOI: 10.1002/jrs.954
    33. 33
      Lee, Y. J.; Grey, C. P. 6Li Magic-Angle Spinning (MAS) NMR Study of Electron Correlations, Magnetic Ordering, and Stability of Lithium Manganese(III) Oxides. Chem. Mater. 2000, 12, 38713878,  DOI: 10.1021/cm000469t
    34. 34
      Hattori, M.; Yamamoto, K.; Matsui, M.; Nakanishi, K.; Mandai, T.; Choudhary, A.; Tateyama, Y.; Sodeyama, K.; Uchiyama, T.; Orikasa, Y.; Tamenori, Y.; Takeguchi, T.; Kanamura, K.; Uchimoto, Y. Role of Coordination Structure of Magnesium Ions on Charge and Discharge Behavior of Magnesium Alloy Electrode. J. Phys. Chem. C 2018, 122, 2520425210,  DOI: 10.1021/acs.jpcc.8b08558
    35. 35
      Yamamoto, K.; Zhou, Y.; Yabuuchi, N.; Nakanishi, K.; Yoshinari, T.; Kobayashi, T.; Kobayashi, Y.; Yamamoto, R.; Watanabe, A.; Orikasa, Y.; Tsuruta, K.; Park, J.; Byon, H. R.; Tamenori, Y.; Ohta, T.; Uchimoto, Y. Charge Compensation Mechanism of Lithium-Excess Metal Oxides with Different Covalent and Ionic Characters Revealed by Operando Soft and Hard X-ray Absorption Spectroscopy. Chem. Mater. 2020, 32, 139147,  DOI: 10.1021/acs.chemmater.9b02838
    36. 36
      Okubo, M.; Yamada, A. Molecular Orbital Principles of Oxygen-Redox Battery Electrodes. ACS Appl. Mater. Interfaces 2017, 9, 3646336472,  DOI: 10.1021/acsami.7b09835
    37. 37
      Yoon, W.-S.; Chung, K. Y.; McBreen, J.; Zaghib, K.; Yang, X.-Q. Electronic Structure of the Electrochemically Delithiated Li[sub 1–x]FePO[sub 4] Electrodes Investigated by P K-edge X-Ray Absorption Spectroscopy. Electrochem. Solid-State Lett. 2006, 9, A415,  DOI: 10.1149/1.2216619
    38. 38
      Bonnet-Mercier, N.; Wong, R. A.; Thomas, M. L.; Dutta, A.; Yamanaka, K.; Yogi, C.; Ohta, T.; Byon, H. R. A structured three-dimensional polymer electrolyte with enlarged active reaction zone for Li–O2 batteries. Sci. Rep. 2015, 4, 7127,  DOI: 10.1038/srep07127
    39. 39
      Rana, J.; Papp, J. K.; Lebens-Higgins, Z.; Zuba, M.; Kaufman, L. A.; Goel, A.; Schmuch, R.; Winter, M.; Whittingham, M. S.; Yang, W.; McCloskey, B. D.; Piper, L. F. J. Quantifying the Capacity Contributions during Activation of Li2MnO3. ACS Energy Letters 2020, 5, 634641,  DOI: 10.1021/acsenergylett.9b02799
    40. 40
      Wandt, J.; Freiberg, A. T. S.; Ogrodnik, A.; Gasteiger, H. A. Singlet oxygen evolution from layered transition metal oxide cathode materials and its implications for lithium-ion batteries. Mater. Today 2018, 21, 825833,  DOI: 10.1016/j.mattod.2018.03.037
    41. 41
      Hu, S.; Li, Y.; Chen, Y.; Peng, J.; Zhou, T.; Pang, W. K.; Didier, C.; Peterson, V. K.; Wang, H.; Li, Q.; Guo, Z. Insight of a Phase Compatible Surface Coating for Long-Durable Li-Rich Layered Oxide Cathode. Adv. Energy Mater. 2019, 9, 1901795,  DOI: 10.1002/aenm.201901795
    42. 42
      Zhang, S.; Gu, H.; Pan, H.; Yang, S.; Du, W.; Li, X.; Gao, M.; Liu, Y.; Zhu, M.; Ouyang, L.; Jian, D.; Pan, F. A Novel Strategy to Suppress Capacity and Voltage Fading of Li- and Mn-Rich Layered Oxide Cathode Material for Lithium-Ion Batteries. Adv. Energy Mater. 2017, 7, 1601066,  DOI: 10.1002/aenm.201601066
    43. 43
      Hekmatfar, M.; Kazzazi, A.; Eshetu, G. G.; Hasa, I.; Passerini, S. Understanding the Electrode/Electrolyte Interface Layer on the Li-Rich Nickel Manganese Cobalt Layered Oxide Cathode by XPS. ACS Appl. Mater. Interfaces 2019, 11, 4316643179,  DOI: 10.1021/acsami.9b14389
    44. 44
      Doi, T.; Shimizu, Y.; Matsumoto, R.; Hashinokuchi, M.; Inaba, M. Suppression of Mn–Ion-Dissolution of LiNi0.5Mn1.5O4 Electrodes in a Highly Concentrated Electrolyte Solution at Elevated Temperatures. ChemistrySelect 2017, 2, 88248827,  DOI: 10.1002/slct.201701668
    45. 45
      Izumi, F.; Momma, K. Three-Dimensional Visualization in Powder Diffraction. Solid State Phenom. 2007, 130, 1520,  DOI: 10.4028/www.scientific.net/SSP.130.15
    46. 46
      Ohara, K.; Tominaka, S.; Yamada, H.; Takahashi, M.; Yamaguchi, H.; Utsuno, F.; Umeki, T.; Yao, A.; Nakada, K.; Takemoto, M.; Hiroi, S.; Tsuji, N.; Wakihara, T. Time-resolved pair distribution function analysis of disordered materials on beamlines BL04B2 and BL08W at SPring-8. J. Synchrotron Radiat. 2018, 25, 16271633,  DOI: 10.1107/S1600577518011232
    47. 47
      Newville, M. IFEFFIT: interactive XAFS analysis and FEFF fitting. J. Synchrotron Radiat. 2001, 8, 322324,  DOI: 10.1107/S0909049500016964
    48. 48
      Hacene, M.; Anciaux-Sedrakian, A.; Rozanska, X.; Klahr, D.; Guignon, T.; Fleurat-Lessard, P. Accelerating VASP electronic structure calculations using graphic processing units. J. Comput. Chem. 2012, 33, 25812589,  DOI: 10.1002/jcc.23096
    49. 49
      Hutchinson, M.; Widom, M. VASP on a GPU: Application to exact-exchange calculations of the stability of elemental boron. Comput. Phys. Commun. 2012, 183, 14221426,  DOI: 10.1016/j.cpc.2012.02.017
    50. 50
      Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B: Condens. Matter Mater. Phys. 1993, 47, 558561,  DOI: 10.1103/PhysRevB.47.558
    51. 51
      Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B: Condens. Matter Mater. Phys. 1996, 54, 1116911186,  DOI: 10.1103/PhysRevB.54.11169
    52. 52
      Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 38653868,  DOI: 10.1103/PhysRevLett.77.3865
    53. 53
      Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 50, 1795317979,  DOI: 10.1103/PhysRevB.50.17953
    54. 54
      Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 59, 17581775,  DOI: 10.1103/PhysRevB.59.1758
    55. 55
      Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104,  DOI: 10.1063/1.3382344
    56. 56
      Jain, A.; Hautier, G.; Ong, S. P.; Moore, C. J.; Fischer, C. C.; Persson, K. A.; Ceder, G. Formation enthalpies by mixing GGA and GGA + U calculations. Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 84, 045115  DOI: 10.1103/PhysRevB.84.045115
    57. 57
      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: Condens. Matter Mater. Phys. 1998, 57, 15051509,  DOI: 10.1103/PhysRevB.57.1505
    58. 58
      Henkelman, G.; Arnaldsson, A.; Jónsson, H. A fast and robust algorithm for Bader decomposition of charge density. Comput. Mater. Sci. 2006, 36, 354360,  DOI: 10.1016/j.commatsci.2005.04.010
    59. 59
      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
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acscentsci.0c01200.

    • DFT calculation details; structural evolutions of the LiMnO2–Li3PO4 system; STEM images; EDX maps; SEM images; comparison of SXRD patterns; Reitveld analysis of rocksalt LiMnO2 and Li7/6P1/6Mn2/3O2; phase segregation of Li7/6P1/6Mn2/3O2 after heating; analysis of the Mn distrivution; X-ray diffraction patterns, X-ray pair distribution functions; Raman, XAS, and solid-state NMR spectra; charge–discharge and charge curves; charge density differences, and synthesis and electrochemical properties of nanosize rocksalt Li2MnO3 (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.

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