Next Article in Journal
Thermal Management in Lithium-Ion Batteries: Latest Advances and Prospects
Previous Article in Journal
Transition Metal-Based Catalysts Powering Practical Room-Temperature Na-S Batteries: From Advances to Further Perspectives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Engineered Plastic–Crystal Interlayers Enable Li-Rich Cathodes in PVDF-HFP-Based All-Solid-State Polymer Batteries

by
Fei Zhou
1,†,
Jinwei Tan
1,†,
Feixiang Wang
1 and
Meiling Sun
1,2,*
1
School of Chemistry, Chemical Engineering and Life Sciences, Wuhan University of Technology, 122 Luoshi Road, Wuhan 430070, China
2
Hubei Longzhong Laboratory, Wuhan University of Technology Xiangyang Demonstration Zone, Xiangyang 441000, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Batteries 2025, 11(9), 334; https://doi.org/10.3390/batteries11090334
Submission received: 26 July 2025 / Revised: 17 August 2025 / Accepted: 2 September 2025 / Published: 6 September 2025

Abstract

All-solid-state lithium batteries (ASSLBs) employing Li-rich layered oxide (LLO) cathodes are regarded as promising next-generation energy storage systems owing to their outstanding energy density and intrinsic safety. Polymer-in-salt solid electrolytes (PISSEs) offer advantages such as high room-temperature ionic conductivity, enhanced Li anode interfacial compatibility, and low processing costs; however, their practical deployment is hindered by poor oxidative stability especially under high-voltage conditions. In this study, we report the rational design of a bilayer electrolyte architecture featuring an in situ solidified LiClO4-doped succinonitrile (LiClO4–SN) plastic–crystal interlayer between a Li1.2Mn0.6Ni0.2O2 (LMNO) cathode and a poly (vinylidene fluoride-co-hexafluoropropylene) (PVDF-HFP)-based PISSE. This PISSE/SN–LiClO4 configuration exhibits a wide electrochemical stability window up to 4.7 V vs. Li+/Li and delivers a high ionic conductivity of 5.68 × 10−4 S cm−1 at 25 °C. The solidified LiClO4-SN layer serves as an effective physical barrier, shielding the PVDF-HFP matrix from direct interfacial contact with LMNO and thereby suppressing its oxidative decomposition at elevated potentials. As a result, the bilayer polymer-based cells with the LMNO cathode demonstrate an initial discharge capacity of ∼206 mAh g−1 at 0.05 C and exhibit good cycling stability with 85.7% capacity retention after 100 cycles at 0.5 C under a high cut-off voltage of 4.6 V. This work not only provides a promising strategy to enhance the compatibility of PVDF-HFP-based electrolytes with high-voltage cathodes through the facile in situ solidification of plastic interlayers but also promotes the application of LMNO cathode material in high-energy ASSLBs.

Graphical Abstract

1. Introduction

Given the escalating demand for high-energy-density applications, such as electric vehicles, unmanned aerial vehicles, and grid-scale energy storage, there is an urgent imperative to further enhance the specific energy density, safety, and lifespan of Li-ion batteries [1]. However, state-of-the-art commercial cathode materials, including polyanion structure (LiFePO4), layered structure (LiCoO2, LiNixCoyMn1-x-yO2 and LiNi0.8Co0.15Al0.05O2), and spinel structure (LiMn2O4), are approaching their theoretical performance limits (<200 mAh g−1). Li-rich Mn-based layered oxides (LRMOs) have emerged as promising cathodes due to their cost-effectiveness, significantly higher specific discharge capacities exceeding 250 mAh g−1, and elevated operating voltages enabled by a synergistic cationic–anionic redox mechanism [2]. Coupling LRMOs with Li metal anodes further enhances the energy density of the resulting full cells. Despite these merits, practical deployment is hindered by poor cycling life and safety concerns. The high-voltage activation of anionic redox processes often induces electrolyte decomposition and lattice oxygen release, which in turn cause interfacial degradation, including cathode–electrolyte interface (CEI) instability and transition metal dissolution [3,4]. Moreover, uncontrolled lithium dendrite growth during repeated stripping/plating cycles can pierce separators, leading to internal short circuits. These interfacial and structural instabilities, together with the flammability and leakage risks of liquid electrolytes, severely compromise battery lifespan, voltage stability, and safety [5,6].
Solid-state electrolyte (SSE) engineering is widely recognized as one of the most promising strategies to effectively address or mitigate the above challenges [7]. State-of-the-art studies of solid-state LRMO|Li batteries reveal that solid electrolytes effectively expand the voltage stability window, mitigate oxygen release, and suppress transition metal leaching [8,9,10,11,12]. However, most of these studies have relied on sulfide- or halide-based SSEs, which still face severe interfacial challenges at both the cathode and anode. Specifically, the high reactivity of Li-rich cathodes can trigger the decomposition of sulfide or halide electrolytes, leading to the formation of insulating derivatives such as Li2SO4, P–O species, and LiCl-rich interphases. These phases increase interfacial impedance and hinder Li+ transport, ultimately degrading cycling stability. On the anode side, sulfide electrolytes undergo reductive decomposition to form electronically conductive byproducts such as Li2S and Li3P, which promote continuous parasitic reactions and lithium dendrite formation [13,14,15]. Halide electrolytes, though relatively more stable, also experience interfacial degradation and typically require artificial buffer layers to suppress undesirable side reactions and maintain interfacial integrity. Furthermore, the inherent mechanical brittleness of inorganic SSEs limits their ability to accommodate volume changes during cycling, leading to delamination and rapid performance degradation.
To address these multifaceted interfacial challenges, polymer-based SSEs, owing to their intrinsic mechanical flexibility and favorable interfacial compatibility, offer a promising pathway for the design of solid LRMO|Li batteries. Although several gel electrolytes have demonstrated encouraging results, very few pure polymer electrolytes (i.e., without liquid components) have shown satisfactory performance in high-voltage LRMO-based systems [16]. Achieving high ionic conductivity, a wide electrochemical stability window, and stable electrode–electrolyte interfaces under high-voltage operation for polymer all-solid-state lithium batteries (ASSLBs) remains a formidable challenge. Conventional single-component polymer electrolytes typically lack a sufficiently wide electrochemical window to simultaneously ensure compatibility with both high-voltage cathodes and lithium metal anodes. For instance, succinonitrile (SN) exhibits appealing properties, including high ionic conductivity (10−3 S cm−1 at RT), a broad electrochemical window (>5 V vs. Li+/Li), and low cost—making it a promising electrolyte candidate for high-voltage ASSLBs. However, SN undergoes a highly spontaneous chemical reaction with Li metal, leading to the severe degradation of the electrolyte/electrode interface [17,18,19,20]. Guo et al. demonstrated that incorporating SN into a poly(ethylene oxide) (PEO)-based electrolyte can suppress PEO decomposition up to 4.3 V when paired with a LiNi0.6Co0.2Mn0.2O2 cathode [21]. Nevertheless, coupling with Li-rich cathodes requires polymer electrolytes with even higher oxidative stability (>4.5 V). Moreover, PEO-based electrolytes inherently suffer from poor ionic conductivity, which severely limits their applicability in high-voltage Li solid-state batteries [22,23].
In this work, we construct a bilayer electrolyte architecture comprising a LiClO4-doped succinonitrile (LiClO4–SN) plastic–crystal interfacing the Li1.2Mn0.6Ni0.2O2 (LMNO) cathode and a poly(vinylidene fluoride-co-hexafluoropropylene) (PVDF-HFP)-based polymer-in-salt solid electrolyte (PISSE) in contact with the Li anode. This design affords a broad electrochemical stability window up to 4.7 V and forms a stable multi-component interphase on the lithium surface. PISSEs have garnered attention due to their high room-temperature ionic conductivity, which can be tuned by simply increasing the salt concentration [24,25,26]. Previously, our group developed a PVDF-HFP-based PISSE (PVHLi-1.1), which was prepared by a facile solution casting method by dissolving PVDF-HFP and LiTFSI at a mass of 1:1.1 in N,N-dimethylformamide (DMF) [27]. The unique ionic clustering in PVHLi-1.1 enables the formation of a robust solid electrolyte interphase (SEI) on lithium, effectively suppressing dendrite growth and supporting the stable cycling of the LiNi0.5Co0.2Mn0.3O2 cathode under 4.3 V. However, to our best knowledge, the application of Li-rich cathodes in PISSE-based ASSLBs remains largely unexplored, primarily due to the limited oxidative stability of PVHLi-1.1 at voltages > 4.3 V. Achieving compatibility with high-voltage Li-rich cathodes while maintaining high energy density and interfacial stability remains a pressing challenge. By leveraging the advantages of SN-LiClO4 and PISSE, we developed a composite PVHLi-1.1/SN-LiClO4 polymer electrolyte that exhibits a wide stability window, high room-temperature ionic conductivity of 5.68 × 10−4 S cm−1, and a relatively stable CEI. This bilayer system enables the successful operation of high-voltage LMNO in ASSLBs (Figure 1), delivering an initial discharge capacity of ∼206 mAh g−1 at 0.05 C and excellent cycling stability, retaining 85.7% capacity after 100 cycles at 0.5 C under a high cut-off voltage of 4.6 V.

2. Materials and Methods

2.1. Materials Preparation

2.1.1. Preparation of LMNO Cathode Material

LMNO was synthesized using the solid-state reaction method as previously reported [2]. Specifically, the precursors of Li2CO3, MnCO3, and Ni(OH)2 (Aladdin, Shanghai, China) with a stoichiometric ratio of Li:Ni:Mn = 1.2:0.2:0.6 were first thoroughly mixed using a planetary ball mill at 400 rpm for 12 h, followed by a pre-annealing process at a temperature of 450 °C for 3 h and a further calcination at 950 °C for 15 h in air.

2.1.2. Preparation of Composite Cathodes with SN-Based Modified Layers

To prepare the composite cathode, LMNO, conductive carbon black (Super P, Shenzhen Kejing Zhida Technology Co., Ltd., Shenzhen, China), and poly(vinylidene fluoride) (PVDF, Sinopharm, Shanghai, China) were mixed in a mass ratio of 80:10:10 in N-methylpyrrolidone (NMP, Sinopharm, Shanghai, China) solvent to form a homogeneous slurry. The slurry was uniformly cast onto aluminum foil and dried in a vacuum oven at 80 °C for over 24 h to obtain the cathode film. For the modification layer, LiClO4 (Aladdin, Shanghai, China) and succinonitrile (SN, Aladdin, Shanghai, China) were mixed in a molar ratio of 1:20 and stirred at 80 °C until a transparent and flowable solution was formed. A certain amount of the resulting LiClO4–SN solution was drop-cast onto the prepared LMNO cathode for 30 min to ensure full infiltration into the porous electrode structure.

2.1.3. Preparation of PVHLi-1.1 and Composite Membranes

The PVHLi-1.1 electrolyte membranes were prepared via a conventional solution casting method. Initially, PVDF-HFP (Mw = 400,000, Sigma-Aldrich, St. Louis, MO, USA) and LiTFSI (Suzhou DuoDuo Chemical Reagent Co., Ltd., Suzhou, China) were dissolved in N,N-Dimethylformamide (DMF, Aladdin, Shanghai, China) at a weight ratio of 10:11:50. The mixture was then mechanically stirred for 5 h at 60 °C to obtain a homogeneous electrolyte slurry. Subsequently, the resulting viscous solution was cast onto a clean glass substrate using a doctor blade. Most of the DMF solvent was removed by air drying under ambient conditions with gentle electrical blowing. Finally, the membranes were further dried at 60 °C in a vacuum oven for 24 h to remove any residual solvent. The resulting PVDF-HFP electrolyte membrane, approximately 100 µm in thickness, was obtained by punching the formed film into circles with a diameter of 16 mm.
Composite membranes were prepared by drop-casting molten SN–LiClO4 onto PVHLi-1.1 placed on a heating plate at 80 °C, followed by cooling to induce solidification.

2.2. Materials Characterization

Powder X-ray diffraction (XRD, Ultima IV, Rigaku, Tokyo, Japan) patterns of the LMNO samples were recorded using a Rigaku Ultima IV diffractometer with Cu Kα radiation (λ = 1.54178 Å) at room temperature. The surface morphology of the composite electrodes was characterized by field emission scanning electron microscopy (SEM, TESCAN MIRA LMS, Brno, Czech Republic), and the corresponding elemental distribution was analyzed using energy-dispersive X-ray spectroscopy (EDS) coupled with SEM. An X-ray photoelectron spectroscopy system (XPS, Thermo Scientific K-Alpha, Waltham, MA, USA) equipped with a monochromatic Al Kα X-ray source was employed to examine the elemental chemical states. High-resolution transmission electron microscopy (HRTEM, Tecnai G2 F20, 200 kV, FEI, Hillsboro, OR, USA) was used to investigate the microstructure of LMNO. Chemical bonding information was collected by Fourier-transform infrared spectroscopy (FTIR, Nicolet 6700, Thermo Fisher Scientific, Waltham, MA, USA), and Raman spectra were acquired on an HR800 spectrometer using a 532 nm excitation laser. For XRD, DSC (NETZSCH-Gerätebau GmbH, Selb, Germany), FTIR, and Raman analyses, the LiClO4–SN solid polymer electrolyte (SPE) samples were prepared by casting the LiClO4–SN solution onto a soft substrate at 80 °C, followed by peeling prior to testing.

2.3. Electrochemical Measurement

2.3.1. Ion Conductivity Testing

The ionic conductivities of the solid polymer electrolyte (SPE) were determined through the electrochemical impedance spectra (EIS) of the stainless-steel SS/SPE/SS blocking cell in the frequency range from 1 HZ to 106 Hz with an alternating current (AC) amplitude of 10 mV at various temperatures from 25 to 80 °C. The ionic conductivity (σ) was calculated using the following equation:
σ = L S R
where R (Ω) is the resistant value of the SPE obtained from the EIS curve, L (cm) denotes the thickness of SPE, and S (cm2) indicates the area of SPE. The activation energy was determined using the Arrhenius Equation:
σ ( T ) = A   exp ( E a RT )
where A is the pre-exponential factor, E a signifies the activation energy of activated ion-hopping conduction process, and T denotes the absolute temperature.

2.3.2. Linear Sweep Voltammetry Testing

Linear sweep voltammetry (LSV) was performed to evaluate the electrochemical stability of the electrolyte, utilizing a scan rate of 1.0 mV s−1 over a voltage range from 2.0 to 6.0 V Li+/Li.

2.3.3. Battery Performance Evaluation

All assembled ASSLBs were cycled between 2.0 and 4.6 V vs. Li+/Li using a LAND CT2001A battery tester (Wuhan LAND Electronics, Wuhan, China) at 25 °C. Charge–discharge stability was assessed under ambient conditions. To further evaluate the rate capability, galvanostatic cycling was conducted at various current rates ranging from 0.05 C to 0.5 C.

3. Results and Discussion

3.1. Characterization of LMNO Cathode

The crystal structure of LMNO was determined by X-ray diffraction (XRD) followed by Rietveld’s refinement. As shown in Figure 2a, all diffraction peaks can be well indexed to the monoclinic phase with the C2/m space group, which also generates several superlattice reflections between 20° and 25°. The refinement parameters are summarized in Table 1. The SEM image (Figure 2b) reveals that the LMNO particles are uniformly distributed, with slight agglomeration, smooth surfaces, and an average size of approximately 250 nm. Further insight into the morphology and crystal structure was provided by high-resolution transmission electron microscopy (HRTEM), as shown in Figure 2c,d. The lattice spacing of 0.472 nm, determined via Fast Fourier Transform (FFT), corresponds well to the (003) planes of the C2/m layered structure. Elemental mapping by SEM-EDX (Figure 2e–g and Figure S1) confirms the uniform distribution of O, Mn, and Ni within the LMNO particles.

3.2. Characterizations of PVDF-HFP PISSE and SN-LiClO4

Figure S3 and Figure 3a–c presents the morphology and intrinsic electrochemical properties of PVHLi-1.1. LiTFSI is uniformly distributed within the PVDF-HFP matrix, enabling a high Li+ transport number of 0.51, and high compatibility with the Li anode. However, although the PVHLi-1.1 electrolyte exhibits a broader electrochemical stability window compared with a normal liquid electrolyte (1 M LiPF6 in EC/DEC) and standard SPEs [24], it remains insufficient to meet the charging demands of the high-voltage LMNO cathode. Figure S2 shows the charge/discharge curve when PVHLi-1.1 is used alone as the SPE, indicating electrolyte decomposition during charging.
To address this, we constructed an in situ cured SN plastic crystal layer doped with LiClO4 between the PVHLi-1.1 electrolyte and the LMNO cathode, leveraging the high oxidative stability of SN and the high conductivity of LiClO4. When the SN plastic crystal was mixed with a defined amount of LiClO4, the resulting blend formed a uniform, flowable, and transparent liquid at 80 °C. As shown in Figure 3d, the mixture remained liquid at 80 °C but solidified upon cooling to room temperature [28]. The XRD pattern (Figure 3e) of the SN–LiClO4 mixture was nearly identical to that of pure SN, indicating complete salt dissociation without disrupting the SN lattice. DSC (Figure 3f) also confirmed the phase transition of SN–LiClO4 from a plastic crystalline to a molten state between 30 and 80 °C. Additionally, Figure S4 displays the magnified FTIR spectra of pure SN and the mixture of SN and LiClO4. The appearance of a characteristic peak at 2281 cm−1 indicates SN solvation and strong SN–Li+ interactions [29,30]. Peaks at 762, 819, 1002, and 1425 cm−1 are attributed to the CH2 group, while the 2254 cm−1 peak corresponds to the C≡N group, consistent with prior reports [31]. Raman spectroscopy further confirmed the complete dissolution of LiClO4 in the SN matrix (Figure 3g).
The calculated activation energies of the PVHLi-1.1/SN–LiClO4 composite were 0.34 eV (low-temperature region) and 0.16 eV (high-temperature region). Notably, the composite achieved a high ionic conductivity of 5.68 × 10−4 S cm−1 at room temperature, nearly eight times higher than that of PVHLi-1.1 alone. (Figure 3h). More importantly, its electrochemical stability window, determined by LSV (Figure 3i), was significantly extended compared with pure PVHLi-1.1. The oxidation onset of PVHLi-1.1 occurs around 4.3 V, intensifying above 4.5 V, thus rendering it incompatible with high-voltage Li-rich cathodes. In contrast, the PVHLi-1.1/SN–LiClO4 electrolyte, benefitting from the oxidative stability of SN’s terminal nitrile groups [32,33], exhibited an expanded potential window up to 4.8 V, making it more compatible with LMNO cathodes. The onset oxidation voltage was determined using the threshold current density method, with the threshold value set at 0.0125 mA cm−2.

3.3. High-Voltage S-LMNO|PVHLi-1.1-SN–LiClO4|Li ASSLB Performance Test

In order to further reduce the interfacial resistance and reduce the side reaction between SN and the lithium anode, the SN-LiClO4 solution was first injected into the porous cathode to form an integrated composite electrode (S-LMNO), and then assembled with PVHLi-1.1 to construct a high-voltage battery to evaluate the electrochemical performance of the intermediate layer. Figure 4d displayed the rate performance of the S-LMNO/PVHLi-1.1/Li cell from 0.05 C to 0.5 C. At a charge cut-off voltage of 4.6 V, the battery delivered discharge specific capacities of approximately 206.3, 195.4, 147.1, and 115.1 mAh g−1 at rates of 0.05 C, 0.1 C, 0.2 C, and 0.5 C, respectively. When the current was changed back to 0.05 C, the discharge capacity recovered to 204.5 mAh g−1. Figure 4a–c and Figure S4 present the charge–discharge performance of the cell at various rates. The good rate performance indicates that the SN intermediate layer possesses good electrical conductivity as well as high oxidation resistance. In terms of cycling performance, after 100 cycles at 0.1, 0.2, and 0.5 C rates, all batteries exhibited capacity retention rates exceeding 84% at 25 °C, achieving excellent cycling stability (Figure 4e,f and Figure S5). Like other reported LRMOs, all cells exhibited a long plateau at ~4.55 V in the high-voltage region, often referred to as the electrochemical activation process. The discharge capacity was close to 200.0 mAh g−1 at 0.1 C, which is slightly lower than the PVDF-HFP-based gel electrolyte [34] but represents a breakthrough for Li-rich manganese-based cathodes in polymer-based ASSLBs. This could be attributed to the good interfacial compatibility and excellent Li+ transfer between the high-voltage LMNO cathode and the PVHLi-1.1-based electrolyte. Additionally, for all batteries tested at different rates, the corresponding high Coulombic efficiency remained above 99%, indicating that the SN-LiClO4 elastomeric interlayer suppressed the decomposition of PVHLi-1.1 at high charging potentials.
Meanwhile, we have also assembled S-LMNO|PVHLi-1.1|Li pouch cells to verify their practical applicability (Figure 5a). By increasing the cathode mass loading, we achieved a breakthrough in energy density. With an LMNO loading of 4.25 mg/cm2, a discharge capacity of 0.72 mAh/cm2 was successfully achieved. The capacity retention was 80.5% even after 40 cycles at a current density of 0.5 C (Figure 5b). As shown in Figure 5c–e, the open-circuit voltage of the assembled ASSLBs did not exhibit a significant decrease during the folding and cutting tests. Figure 5f–h depicts the abuse experiments on the pouch cell, where it was observed that the cell continued to illuminate a high-intensity LED bulb, even after being subjected to folding and cutting damage. Notably, there were no indications of internal short circuits or combustion, and the brightness of the LED remained virtually unchanged, underscoring the battery’s exceptional safety performance. These tests confirm that the S-LMNO/PVHLi-1.1/Li battery holds promising potential for practical applications.

3.4. Interface Analysis

To gain deeper insights into the role of the SN–LiClO4 interlayer in regulating cathode–electrolyte interphase (CEI) formation in the LMNO|PVHLi-1.1|Li cell, detailed ex situ XPS measurements were conducted to collect O 1s spectra from P-LMNO (without the SN–LiClO4 interlayer) and S-LMNO composite cathodes at both the fully charged (C-4.6 V) and fully discharged (D-2.0 V) states after the first cycle.
As shown in Figure 6a,d, for the P-LMNO electrode charged to 4.6 V and discharged to 2.0 V, two dominant O 1s peaks were identified at 532.8 eV (purple) and 531.8 eV (blue), corresponding to surface oxygenated species such as C–O/C=O and LiOH/Li2CO3, respectively [27,31,35]. These species are attributed to CEI formation and intensive parasitic reactions, including the decomposition of TFSI anions during cycling. Notably, the O 1s signal corresponding to the lattice oxygen (O2−, green) in LMNO was absent, likely due to the extensive coverage by decomposition products on the electrode surface. After Ar+ sputtering to a depth of ~20 nm, weak lattice oxygen signals emerged in both electrodes. Additionally, a minor peak at ~530.5 eV (red) appeared in the charged P-LMNO sample (D-2.0 V), which is assigned to low-electron-density oxygen species (O2n−), indicative of surface-localized anionic redox activity or oxidative oxygen intermediates [31,36].
In contrast, the S-LMNO electrode at both the charged and discharged states (Figure 6b,e) exhibited markedly fewer organic byproducts, indicating suppressed electrolyte decomposition. Moreover, more pronounced signals corresponding to lattice oxygen (O2−, ~529.6 eV) and low-electron-density oxygen species (O2n−) were observed, suggesting improved lattice oxygen stability and the enhanced participation of anionic redox processes. Notably, the binding energy of O2n− on the surface shifted to a higher value (~531.0 eV), which likely reflects an evolution of the oxygen electronic structure—from a delocalized high-valence state near the surface to a more localized configuration within the bulk lattice [37,38]. Upon discharging to 2.0 V, the surface O2n− signal in S-LMNO decreased significantly, further confirming that the introduction of the SN–LiClO4 plastic crystal interlayer effectively mitigates interfacial parasitic reactions and promotes the reversibility of the oxygen redox process.
Furthermore, XPS analysis was performed on the surface of the Li anode retrieved from the S-LMNO|PVHLi-1.1|Li cell after 100 cycles, as shown in Figure 7. The interfacial layer mainly consisted of LiF and Li3N, forming a composite solid electrolyte interphase (SEI). Each component plays a distinct yet complementary role in stabilizing the Li/electrolyte interface. Specifically, LiF is an electronically insulating and chemically robust species that effectively blocks electron leakage and suppresses continuous electrolyte decomposition, while its high mechanical strength helps inhibit lithium dendrite penetration [39,40]. Li3N, on the other hand, is a fast Li+ conductor (σ ≈ 10−3 S·cm−1) that facilitates efficient ion transport across the SEI and reduces interfacial resistance [27,41]. By integrating these complementary properties, the LiF–Li3N composite SEI simultaneously provides high Li+ conductivity, superior electronic insulation, enhanced chemical stability, and mechanical robustness. Such a synergistic interphase not only suppresses parasitic side reactions but also promotes uniform lithium deposition, thereby sustaining reversible Li plating/stripping and significantly enhancing the overall electrochemical performance of the cell [42].

4. Conclusions

In summary, to enable high-voltage all-solid-state Li-rich manganese-based oxide (LRMO)|Li batteries, a bilayer polymer electrolyte was constructed by introducing an SN–LiClO4 interlayer in situ between the PVHLi-1.1 PISSE and the high-voltage cathode. This strategy effectively expanded the electrochemical stability window to 4.7 V and enhanced the room-temperature ionic conductivity to 5.68 × 10−4 S cm−1. As a result, the assembled solid-state battery delivered a high initial capacity of 206 mAh g−1 within 2.4–4.6 V, and exhibited an outstanding cycling stability with 85.7% capacity retention after 100 cycles at 0.5 C under a high cut-off voltage of 4.6 V. The solidified SN layer further acted as a robust physical barrier, mitigating the oxidative decomposition of PVDF-HFP induced by the highly reactive LRMO cathode. XPS analysis confirmed that the SN–LiClO4 layer formed in situ effectively suppressed interfacial side reactions and enhanced the reversibility of anionic (oxygen) redox processes. The assembled solid-state cell also exhibited excellent mechanical integrity and safety performance under external mechanical stress. This work provides a promising strategy to advance the practical application of high-energy Li-rich cathodes in all-solid-state lithium batteries.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/batteries11090334/s1, Figure S1: Morphology of PVHLi-1.1 electrolyte; Figure S2: Charge–discharge curves of solid LMNO|Li cell using PVHLi-1.1 PISSE without SN-LiClO4 interlayer at 0.1 C; Figure S3: Morphology and EDS of PVHLi-1.1 electrolyte; Figure S4: FTIR spectra of SN and the SN-LiClO4 interlayer; Figure S5: Charge–discharge performance of the S-LMNO|PVHLi-1.1|Li cell at 0.05C; Figure S6: Cycling performance of th e S-LMNO|PVHLi-1.1|Li cell at 0.2C within a voltage range of 2.0–4.6 V.

Author Contributions

Conceptualization, M.S.; methodology, F.Z. and F.W.; formal analysis, F.Z. and J.T.; writing—original draft preparation, F.Z. and J.T.; writing and supervision—review and editing, M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (22309142).

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Acknowledgments

We would like to thank Jinping Liu and Liang Xiao (Wuhan University of Technologies) for the fruitful discussions.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Du, H.; Zhang, X.; Yu, H. Design of high-energy-density lithium batteries: Liquid to all solid state. eTransportation 2025, 23, 100382. [Google Scholar] [CrossRef]
  2. Xu, J.; Sun, M.; Qiao, R.; Renfrew, S.E.; Ma, L.; Wu, T.; Hwang, S.; Nordlund, D.; Su, D.; Amine, K.; et al. Elucidating Anionic Oxygen Activity in Lithium-rich Layered Oxides. Nat. Commun. 2018, 9, 947. [Google Scholar] [CrossRef]
  3. Wang, H.; Geng, X.; Hu, L.; Wang, J.; Xu, Y.; Zhu, Y.; Liu, Z.; Lu, J.; Lin, Y.; He, X. Efficient Direct Repairing of Lithium- and Manganese-rich Cathodes by Concentrated Solar Radiation. Nat. Commun. 2024, 15, 1634. [Google Scholar] [CrossRef]
  4. Yuan, X.; Dong, T.; Liu, J.; Cui, Y.; Dong, H.; Yuan, D.; Zhang, H. Bi-affinity Electrolyte Optimizing High-Voltage Lithium-Rich Manganese Oxide Battery via Interface Modulation Strategy. Angew. Chem. Int. Ed. 2023, 62, e202304121. [Google Scholar] [CrossRef]
  5. Liu, W.; Li, J.; Li, W.; Xu, H.; Zhang, C.; Qiu, X. Inhibition of Transition Metals Dissolution in Cobalt-free Cathode with Ultrathin Robust Interphase in Concentrated Electrolyte. Nat. Commun. 2020, 11, 3629. [Google Scholar] [CrossRef]
  6. Betz, J.; Brinkmann, J.-P.; Nölle, R.; Lürenbaum, C.; Kolek, M.; Stan, M.C.; Winter, M.; Placke, T. Cross Talk between Transition Metal Cathode and Li Metal Anode: Unraveling Its Influence on the Deposition/Dissolution Behavior and Morphology of Lithium. Adv. Energy Mater. 2019, 9, 1900574. [Google Scholar] [CrossRef]
  7. Antony Jose, S.; Gallant, A.; Gomez, P.L.; Jaggers, Z.; Johansson, E.; LaPierre, Z.; Menezes, P.L. Solid-State Lithium Batteries: Advances, Challenges, and Future Perspectives. Batteries 2025, 11, 90. [Google Scholar] [CrossRef]
  8. Hu, N.; Zhang, Y.-H.; Yang, Y.; Wu, H.; Liu, Y.; Hao, C.; Zheng, Y.; Sun, D.; Li, W.; Li, J.; et al. Unraveling the Spatial Asynchronous Activation Mechanism of Oxygen Redox-Involved Cathode for High-Voltage Solid-State Batteries. Adv. Energy Mater. 2024, 14, 2303797. [Google Scholar] [CrossRef]
  9. Yu, R.; Wang, C.; Duan, H.; Jiang, M.; Zhang, A.; Fraser, A.; Zuo, J.; Wu, Y.; Sun, Y.; Zhao, Y.; et al. Manipulating Charge-Transfer Kinetics of Lithium-Rich Layered Oxide Cathodes in Halide All-Solid-State Batteries. Adv. Mater. 2023, 35, 2207234. [Google Scholar] [CrossRef]
  10. Wu, Y.; Zhou, K.; Ren, F.; Ha, Y.; Liang, Z.; Zheng, X.; Wang, Z.; Yang, W.; Zhang, M.; Luo, M.; et al. Highly Reversible Li2RuO3 Cathodes in Sulfide-based All Solid-state Lithium Batteries. Energy Environ. Sci. 2022, 15, 3470–3482. [Google Scholar] [CrossRef]
  11. Sun, S.; Zhao, C.Z.; Yuan, H.; Fu, Z.H.; Chen, X.; Lu, Y.; Li, Y.F.; Hu, J.K.; Dong, J.; Huang, J.Q.; et al. Eliminating Interfacial O-involving Degradation in Li-rich Mn-based Cathodes for All-solid-state Lithium Batteries. Sci. Adv. 2022, 8, eadd5189. [Google Scholar] [CrossRef]
  12. Liu, B.; Hu, N.; Li, C.; Ma, J.; Zhang, J.; Yang, Y.; Sun, D.; Yin, B.; Cui, G. Direct Observation of Li-Ion Transport Heterogeneity Induced by Nanoscale Phase Separation in Li-rich Cathodes of Solid-State Batteries. Angew. Chem. Int. Ed. 2022, 61, e202209626. [Google Scholar] [CrossRef] [PubMed]
  13. Cao, C.; Carbone, M.R.; Komurcuoglu, C.; Shekhawat, J.S.; Sun, K.; Guo, H.; Liu, S.; Chen, K.; Bak, S.-M.; Du, Y.; et al. Atomic Insights into the Oxidative Degradation Mechanisms of Sulfide Solid Electrolytes. Cell Rep. Phys. Sci. 2024, 5, 101909. [Google Scholar] [CrossRef]
  14. Li, J.; Luo, J.; Li, X.; Fu, Y.; Zhu, J.; Zhuang, X. Li Metal Anode Interface in Sulfide-based All-solid-state Li batteries. EcoMat 2023, 5, e12383. [Google Scholar] [CrossRef]
  15. Byeon, Y.-W.; Kim, H. Review on Interface and Interphase Issues in Sulfide Solid-State Electrolytes for All-Solid-State Li-Metal Batteries. Electrochem 2021, 2, 452–471. [Google Scholar] [CrossRef]
  16. Wang, H.; Yang, Y.; Gao, C.; Chen, T.; Song, J.; Zuo, Y.; Fang, Q.; Yang, T.; Xiao, W.; Zhang, K.; et al. An Entanglement Association Polymer electrolyte for Li-metal batteries. Nat. Commun. 2024, 15, 2500. [Google Scholar] [CrossRef]
  17. Chen, S.; Wang, S.; Peng, Q.; Wei, Z.; Cheng, S.; Fang, Z.; Duan, P.; Cheng, Y.; Cheng, Y.; Jin, K.; et al. In-Situ Fabricated Succinonitrile-based Composite Electrolyte for High-performance and Safe Solid-state Lithium Batteries. J. Power Sources 2024, 604, 234473. [Google Scholar] [CrossRef]
  18. Chen, J.; Yang, Z.; Xu, X.; Qiao, Y.; Zhou, Z.; Hao, Z.; Chen, X.; Liu, Y.; Wu, X.; Zhou, X.; et al. Nonflammable Succinonitrile-Based Deep Eutectic Electrolyte for Intrinsically Safe High-Voltage Sodium-Ion Batteries. Adv. Mater. 2024, 36, 2400169. [Google Scholar] [CrossRef]
  19. Bao, D.; Tao, Y.; Zhong, Y.; Zhao, W.; Peng, M.; Zhang, H.; Sun, X. High-Performance Dual-Salt Plastic Crystal Electrolyte Enabled by Succinonitrile-Regulated Porous Polymer Host. Adv. Funct. Mater. 2023, 33, 2213211. [Google Scholar] [CrossRef]
  20. Das, S.; Prathapa, S.J.; Menezes, P.V.; Row, T.N.G.; Bhattacharyya, A.J. Study of Ion Transport in Lithium Perchlorate-Succinonitrile Plastic Crystalline Electrolyte via Ionic Conductivity and In Situ Cryo-Crystallography. J. Phys. Chem. B 2009, 113, 5025–5031. [Google Scholar] [CrossRef] [PubMed]
  21. Zuo, M.; Bi, Z.; Guo, X. In-Situ solidification of plastic interlayers enabling high-voltage solid lithium batteries with poly(ethylene oxide) based polymer electrolytes. Chem. Eng. J. 2023, 463, 142463. [Google Scholar] [CrossRef]
  22. Naboulsi, A.; Chometon, R.; Ribot, F.; Nguyen, G.; Fichet, O.; Laberty-Robert, C. Correlation between Ionic Conductivity and Mechanical Properties of Solid-like PEO-based Polymer Electrolyte. ACS Appl. Mater. Interfaces 2024, 16, 13869–13881. [Google Scholar] [CrossRef]
  23. An, Y.; Han, X.; Liu, Y.; Azhar, A.; Na, J.; Nanjundan, A.K.; Wang, S.; Yu, J.; Yamauchi, Y. Progress in Solid Polymer Electrolytes for Lithium-Ion Batteries and Beyond. Small 2021, 18, 2103617. [Google Scholar] [CrossRef]
  24. Li, Y.; Yuan, W.; Hu, Z.; Shen, Y.; Wu, G.; Cong, F.; Fu, X.; Lu, F.; Li, Y.; Liu, P.; et al. Constructing PVDF-Based Polymer Electrolyte for Lithium Metal Batteries by Polymer-Induced Phase Structure Adjustment Strategy. Adv. Funct. Mater. 2025, 35, 2424763. [Google Scholar] [CrossRef]
  25. Xiong, Z.; Wang, Z.; Zhou, W.; Liu, Q.; Wu, J.-F.; Liu, T.-H.; Xu, C.; Liu, J. 4.2V Polymer All-solid-state Lithium Batteries Enabled by High-concentration PEO Solid Electrolytes. Energy Storage Mater. 2023, 57, 171–179. [Google Scholar] [CrossRef]
  26. Panneerselvam, T.; Murugan, R.; Sreejith, O.V.; Moulisvar, M.; Elsin Abraham, S. 3D Flexible Electrospun Nanocomposite Polymer Electrolyte Based on Li1.45Al0.45Ge0.2Ti1.35(PO4)3 for Lithium Metal Batteries. Energy Fuels 2023, 38, 682–693. [Google Scholar] [CrossRef]
  27. Liu, W.; Yi, C.; Li, L.; Liu, S.; Gui, Q.; Ba, D.; Li, Y.; Peng, D.; Liu, J. Designing Polymer-in-Salt Electrolyte and Fully Infiltrated 3D Electrode for Integrated Solid-State Lithium Batteries. Angew. Chem. Int. Ed. 2021, 60, 12931–12940. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, Q.; Yu, Q.; Li, S.; Wang, S.; Zhang, L.; Cai, B.; Zhou, D.; Li, B. Safe LAGP-based All Solid-state Li Metal Batteries with Plastic Super-conductive Interlayer Enabled by In-Situ Solidification. Energy Storage Mater. 2020, 25, 613–620. [Google Scholar] [CrossRef]
  29. Hu, Z.; Xian, F.; Guo, Z.; Lu, C.; Du, X.; Cheng, X.; Zhang, S.; Dong, S.; Cui, G.; Chen, L. Nonflammable Nitrile Deep Eutectic Electrolyte Enables High-Voltage Lithium Metal Batteries. Chem. Mater. 2020, 32, 3405–3413. [Google Scholar] [CrossRef]
  30. Wang, Q.; Fan, H.; Fan, L.Z.; Shi, Q. Preparation and Performance of a Non-ionic Plastic Crystal Electrolyte with the Addition of Polymer for Lithium Ion Batteries. Electrochim. Acta 2013, 114, 720–725. [Google Scholar] [CrossRef]
  31. Shen, C.; Liu, Y.; Hu, L.; Li, W.; Liu, X.; Shi, Y.; Jiang, Y.; Zhao, B.; Zhang, J. Regulating Anionic Redox Activity of Lithium-rich Layered Oxides via LiNbO3 Integrated Modification. Nano Energy 2022, 101, 107555. [Google Scholar] [CrossRef]
  32. Zhao, B.; Yang, M.; Li, J.; Li, S.; Zhang, G.; Liu, S.; Cui, Y.; Liu, H. Cellulose-Based Plastic Crystal Electrolyte Membranes with Enhanced Interface for Solid-State Lithium Batteries. Energy Technol. 2021, 9, 2100114. [Google Scholar] [CrossRef]
  33. Arunkumar, R.; Babu, R.S.; Usha Rani, M.; Kalainathan, S. Effect of PBMA on PVC-based Polymer Blend Electrolytes. J. Appl. Polym. Sci. 2017, 134, 44939. [Google Scholar] [CrossRef]
  34. Bian, X.; Liang, J.; Tang, X.; Li, R.; Kang, L.; Su, A.; Su, X.; Wei, Y. A Boron Nitride-polyvinylidene Fluoride-co-hexafluoropropylene Composite Gel Polymer Electrolyte for Lithium Metal Batteries. J. Alloys Compd. 2019, 803, 1075–1081. [Google Scholar] [CrossRef]
  35. Zhang, L.; Xu, X.; Jiang, S.; Wei, L.; Xi, K.; Lei, Y.; Cheng, X.; Yin, J.; Gao, Y. Halloysite Nanotubes Modified Poly(vinylidenefluoride-co-hexafluoropropylene)-based Polymer-in-salt Electrolyte to Achieve High-performance Li Metal Batteries. J. Colloid Interface Sci. 2023, 645, 45–54. [Google Scholar] [CrossRef] [PubMed]
  36. Sathiya, M.; Rousse, G.; Ramesha, K.; Laisa, C.P.; Vezin, H.; Sougrati, M.T.; Doublet, M.L.; Foix, D.; Gonbeau, D.; Walker, W. Reversible Anionic Redox Chemistry in High-capacity Layered-oxide Electrodes. Nat. Mater. 2013, 12, 827–835. [Google Scholar] [CrossRef]
  37. Naylor, A.J.; Makkos, E.; Maibach, J.; Guerrini, N.; Sobkowiak, A.; Björklund, E.; Lozano, J.G.; Menon, A.S.; Younesi, R.; Roberts, M.R. Depth-dependent Oxygen Redox Activity in Lithium-rich Layered Oxide Cathodes. J. Mater. Chem. A 2019, 7, 25355–25368. [Google Scholar] [CrossRef]
  38. Shimoda, K.; Minato, T.; Nakanishi, K.; Komatsu, H.; Matsunaga, T.; Tanida, H.; Arai, H.; Ukyo, Y.; Uchimoto, Y.; Ogumi, Z. Oxidation Behavior of Lattice Oxygen in Li-rich Manganese-based Layered Oxide Studied by Hard X-ray Photoelectron Spectroscopy. J. Mater. Chem. A 2016, 4, 5909–5916. [Google Scholar] [CrossRef]
  39. Liang, W.; Zhou, X.; Zhang, B.; Zhao, Z.; Song, X.; Chen, K.; Wang, L.; Ma, Z.; Liu, J. The Versatile Establishment of Charge Storage in Polymer Solid Electrolyte with Enhanced Charge Transfer for LiF-Rich SEI Generation in Lithium Metal Batteries. Angew. Chem. Int. Ed. 2024, 63, e202320149. [Google Scholar] [CrossRef] [PubMed]
  40. Zheng, Y.; Yang, N.; Duan, S.; Li, Z.; Gao, R.; Zhu, Y.; Wang, H.; Zhang, T.; Li, G.; Luo, D.; et al. Dual-Enhanced Charge Transfer through Prelithiation Strategy in Polymer Electrolyte Enables Robust LiF-Rich SEI for Ultralong-Life All-Solid-State Batteries. Adv. Funct. Mater. 2025, e11011. [Google Scholar] [CrossRef]
  41. Kim, M.S.; Zhang, Z.; Wang, J.; Oyakhire, S.T.; Kim, S.C.; Yu, Z.; Chen, Y.; Boyle, D.T.; Ye, Y.; Huang, Z.; et al. Revealing the Multifunctions of Li3N in the Suspension Electrolyte for Lithium Metal Batteries. ACS Nano 2023, 17, 3168–3180. [Google Scholar] [CrossRef] [PubMed]
  42. Xie, X.; Wang, Z.; He, S.; Chen, K.; Huang, Q.; Zhang, P.; Hao, S.M.; Wang, J.; Zhou, W. Influencing Factors on Li-ion Conductivity and Interfacial Stability of Solid Polymer Electrolytes, Exampled by Polycarbonates, Polyoxalates and Polymalonates. Angew. Chem. Int. Ed. 2023, 62, e202218229. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration of the ASSLB configuration consisting of an LMNO cathode and a PVHLi-1.1 electrolyte, highlighting the interface modification achieved via the introduction of an SN-LiClO4 interlayer. The PVHLi-1.1 PISSE provides high ionic conductivity and forms a robust SEI on the Li anode, while the SN-LiClO4 interlayer further stabilizes the cathode–electrolyte interface and suppresses the decomposition of the solid electrolyte.
Figure 1. Schematic illustration of the ASSLB configuration consisting of an LMNO cathode and a PVHLi-1.1 electrolyte, highlighting the interface modification achieved via the introduction of an SN-LiClO4 interlayer. The PVHLi-1.1 PISSE provides high ionic conductivity and forms a robust SEI on the Li anode, while the SN-LiClO4 interlayer further stabilizes the cathode–electrolyte interface and suppresses the decomposition of the solid electrolyte.
Batteries 11 00334 g001
Figure 2. (a) XRD Rietveld’s refinement of LMNO sample; (b) SEM images of LMNO sample; (c,d) HRTEM images of LMNO sample and corresponding Fast Fourier Transform (FFT); (eg) EDS mapping (O, Mn, and Ni) images of LMNO sample.
Figure 2. (a) XRD Rietveld’s refinement of LMNO sample; (b) SEM images of LMNO sample; (c,d) HRTEM images of LMNO sample and corresponding Fast Fourier Transform (FFT); (eg) EDS mapping (O, Mn, and Ni) images of LMNO sample.
Batteries 11 00334 g002
Figure 3. (ac) Characterization of the PVHLi-1.1 electrolyte: (a) Arrhenius plot of ionic conductivity; (b) polarization curves of PVHLi-1.1 before and after polarization at 25 °C; (c) voltage profile of a symmetric Li/PVHLi-1.1/Li cell at a current density of 0.1 mA cm−2. (dg) Properties of the SN-LiClO4: (d) photographs of the SN-LiClO4 electrolyte in its liquid and solid states at 80 °C and 30 °C, respectively; (e) XRD patterns of LiClO4 powder, pure SN, and the SN-LiClO4 interlayer; (f) DSC curve of the SN-LiClO4 electrolyte at a heating rate of 10 °C min−1; (g) Raman spectra of LiClO4 powder, SN, and SN-LiClO4 layer. (h,i) Performance of the PVHLi-1.1/SN-LiClO4 composite electrolyte: (h) ionic conductivity of the PVHLi-1.1/SN-LiClO4 composite compared with that of PVHLi-1.1. (i) LSV curves of the PVHLi-1.1/SN-LiClO4 composite and PVHLi-1.1.
Figure 3. (ac) Characterization of the PVHLi-1.1 electrolyte: (a) Arrhenius plot of ionic conductivity; (b) polarization curves of PVHLi-1.1 before and after polarization at 25 °C; (c) voltage profile of a symmetric Li/PVHLi-1.1/Li cell at a current density of 0.1 mA cm−2. (dg) Properties of the SN-LiClO4: (d) photographs of the SN-LiClO4 electrolyte in its liquid and solid states at 80 °C and 30 °C, respectively; (e) XRD patterns of LiClO4 powder, pure SN, and the SN-LiClO4 interlayer; (f) DSC curve of the SN-LiClO4 electrolyte at a heating rate of 10 °C min−1; (g) Raman spectra of LiClO4 powder, SN, and SN-LiClO4 layer. (h,i) Performance of the PVHLi-1.1/SN-LiClO4 composite electrolyte: (h) ionic conductivity of the PVHLi-1.1/SN-LiClO4 composite compared with that of PVHLi-1.1. (i) LSV curves of the PVHLi-1.1/SN-LiClO4 composite and PVHLi-1.1.
Batteries 11 00334 g003
Figure 4. (ac) Charge–discharge profiles at different cut-off voltages under current densities of (a) 0.1 C, (b) 0.2 C, and (c) 0.5 C. (d) Rate capability of the S-LMNO/PVHLi-1.1/Li cell at various cut-off voltages under current densities ranging from 0.05 C to 0.5 C. (e,f) Cycling performance and Coulombic efficiency at (e) 0.1 C and (f) 0.5 C, respectively. All electrochemical measurements were conducted at 25 °C.
Figure 4. (ac) Charge–discharge profiles at different cut-off voltages under current densities of (a) 0.1 C, (b) 0.2 C, and (c) 0.5 C. (d) Rate capability of the S-LMNO/PVHLi-1.1/Li cell at various cut-off voltages under current densities ranging from 0.05 C to 0.5 C. (e,f) Cycling performance and Coulombic efficiency at (e) 0.1 C and (f) 0.5 C, respectively. All electrochemical measurements were conducted at 25 °C.
Batteries 11 00334 g004
Figure 5. (a) Schematic illustration of the pouch cell configuration; (b) cycling performance of the S-LMNO/PVHLi-1.1/Li cell, with a cathode loading of 4.25 mg/cm2 and a solid electrolyte thickness of ~100 μm; (ce) open-circuit voltage of the pouch cell in different states; (fh) high-voltage Li-rich/Li pouch cell lighting a LED under extreme damage conditions.
Figure 5. (a) Schematic illustration of the pouch cell configuration; (b) cycling performance of the S-LMNO/PVHLi-1.1/Li cell, with a cathode loading of 4.25 mg/cm2 and a solid electrolyte thickness of ~100 μm; (ce) open-circuit voltage of the pouch cell in different states; (fh) high-voltage Li-rich/Li pouch cell lighting a LED under extreme damage conditions.
Batteries 11 00334 g005
Figure 6. (a,d) O 1s XPS spectra of the ex situ P-LMNO cathodes cycled to 4.6 V and 2.0 V during the initial cycle; (b,e) O 1s XPS spectra of the ex situ S-LMNO cathodes cycled to 4.6 V and 2.0 V during the initial cycle; (c,f) composition analysis of different oxide species at different depths of the LMNO cathodes corresponding to (a,b) and (d,e).
Figure 6. (a,d) O 1s XPS spectra of the ex situ P-LMNO cathodes cycled to 4.6 V and 2.0 V during the initial cycle; (b,e) O 1s XPS spectra of the ex situ S-LMNO cathodes cycled to 4.6 V and 2.0 V during the initial cycle; (c,f) composition analysis of different oxide species at different depths of the LMNO cathodes corresponding to (a,b) and (d,e).
Batteries 11 00334 g006
Figure 7. (a) F 1s, (b) N 1s, and (c) C 1s XPS spectra of the Li anode retrieved from the S-LMNO|PVHLi-1.1|Li cell after 100 cycles.
Figure 7. (a) F 1s, (b) N 1s, and (c) C 1s XPS spectra of the Li anode retrieved from the S-LMNO|PVHLi-1.1|Li cell after 100 cycles.
Batteries 11 00334 g007
Table 1. Structural solution of LMNO sample determined by XRD Rietveld’s refinement.
Table 1. Structural solution of LMNO sample determined by XRD Rietveld’s refinement.
Li1.2Mn0.6Ni0.2O2, C2/m, a = 4.954(6) Å, b = 8.563(3) Å, c = 5.030(6) Å, V = 201.4(2),
β = 109.25(1)°, Rwp = 3.124%
Atom Position OccupancyUisoMultiplicity
Li100.1607(11)00.30.029(3)4
Mn100.1607(11)00.70.029(3)4
Ni100.500.60.012(4)2
Mn200.500.40.012(4)2
Li3000.510.0352
Li400.666020.510.0354
O10.238(4)00.257(3)10.013(2)4
O20.232(3)0.346(1)0.213(1)10.013(1)8
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, F.; Tan, J.; Wang, F.; Sun, M. In Situ Engineered Plastic–Crystal Interlayers Enable Li-Rich Cathodes in PVDF-HFP-Based All-Solid-State Polymer Batteries. Batteries 2025, 11, 334. https://doi.org/10.3390/batteries11090334

AMA Style

Zhou F, Tan J, Wang F, Sun M. In Situ Engineered Plastic–Crystal Interlayers Enable Li-Rich Cathodes in PVDF-HFP-Based All-Solid-State Polymer Batteries. Batteries. 2025; 11(9):334. https://doi.org/10.3390/batteries11090334

Chicago/Turabian Style

Zhou, Fei, Jinwei Tan, Feixiang Wang, and Meiling Sun. 2025. "In Situ Engineered Plastic–Crystal Interlayers Enable Li-Rich Cathodes in PVDF-HFP-Based All-Solid-State Polymer Batteries" Batteries 11, no. 9: 334. https://doi.org/10.3390/batteries11090334

APA Style

Zhou, F., Tan, J., Wang, F., & Sun, M. (2025). In Situ Engineered Plastic–Crystal Interlayers Enable Li-Rich Cathodes in PVDF-HFP-Based All-Solid-State Polymer Batteries. Batteries, 11(9), 334. https://doi.org/10.3390/batteries11090334

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop