Zhongxiu Liuabc,
Md Shariful Islamd,
Yuhui Fange,
Meifang Zhua,
Changyong (Chase) Cao
*df and
Guiyin Xu
*a
aState Key Laboratory for Modification of Chemical Fibers and Polymer Materials, College of Materials Science and Engineering, Donghua University, Shanghai 201620, China. E-mail: xuguiyin@dhu.edu.cn
bHenan Academy of Sciences, Zhengzhou 450001, China
cSchool of Materials Science and Engineering, Zhengzhou University, Zhengzhou 450001, China
dLaboratory for Soft Machines and Electronics, Department of Mechanical and Aerospace Engineering, Case Western Reserve University, Cleveland, OH 44106, USA. E-mail: ccao@case.edu
e4D Maker LLC, Okemos, MI 44106, USA
fAdvanced Platform Technology (APT) Center, Louis Stokes Cleveland VA Medical Center, Cleveland, OH 44106, USA
First published on 26th December 2024
Lithium metal is considered one of the most promising anode materials for lithium batteries due to its high theoretical specific capacity (3860 mA h g−1) and low redox potential (−3.04 V). However, uncontrolled lithium dendrite growth and severe interfacial side reactions during cycling result in poor performance and safety risks, significantly limiting its practical applications. Replacing liquid electrolytes with solid polymer electrolytes (SPEs) offers a solution, as SPEs provide flexibility and good electrode compatibility, effectively inhibiting dendrite growth and reducing interfacial reactions. Among SPEs, poly(vinylidene fluoride) (PVDF)-based solid electrolytes offer excellent thermal stability and mechanical strength, making them highly suitable for high-energy-density flexible batteries. This review presents recent advances in PVDF-based solid-state electrolytes (SSEs) for stable, high-performance lithium metal batteries (LMBs). We focus on modification strategies that enhance the performance of PVDF-based SSEs in solid-state LMBs and highlight how synthesis methods, nano/microstructural design, and electrochemical properties are interrelated. Lastly, we discuss the challenges and prospects for PVDF-based SSEs in next-generation high-performance LMBs.
Compared to commercial lithium-ion batteries, lithium metal batteries (LMBs), which use metallic lithium directly as the anode, offer higher specific energy based on the electroplating and stripping of lithium ions.11 Lithium metal has an exceptionally high theoretical specific capacity (3860 mA h g−1) and a very low reduction potential (−3.04 V),12 making LMBs with lithium metal anodes highly promising for achieving high energy densities. Additionally, lithium metal can be used as the anode material directly, eliminating the need for heavy and inactive current collectors, thus further increasing the battery's specific energy.13 The lithium metal anode (LMA) holds significant potential for development and is highly anticipated as a transformative solution.14 However, several challenges currently limit its practical application.15 Firstly, the volume expansion of the lithium metal anode during cycling causes instability in the solid electrolyte interface (SEI) layer, leading to reduced coulombic efficiency and accelerated capacity decay.16 Secondly, the uncontrolled growth of lithium dendrites can penetrate the separator, causing short circuits, fire hazards, and other safety risks.17
Organic liquid electrolytes, characterized by high volatility, flammability, and leakage risks, pose inherent safety issues for lithium metal batteries.18 In contrast, replacing liquid electrolytes with solid-state electrolytes (SSEs) can address these concerns by significantly improving the safety and energy density of lithium metal batteries.19,20 Solid electrolytes are typically classified as either inorganic or polymer-based.21–23 Inorganic SSEs offer excellent mechanical properties, thermal stability, and high ionic conductivity,24 with primary examples including oxides, sulfides, and halides.25–27 However, large-scale application faces challenges. Although oxide-based SSEs exhibit relatively high ionic conductivity, their rigidity complicates battery processing and assembly and increases interface resistance.28 During preparation, oxide SSEs often require prolonged high-speed ball milling and elevated temperatures to enhance uniformity and density, resulting in high manufacturing costs. Sulfide and halide SSEs, while having ionic conductivities comparable to liquid electrolytes, generally exhibit narrow electrochemical windows, complicating direct compatibility with high-voltage cathode materials.28–31
Polymer-based SSEs, on the other hand, are highly flexible, capable of forming strong interfacial contact with both cathodes and anodes, and exhibit low interfacial impedance, offering significant practical potential.32 Common polymer substrates include poly(ethylene oxide) (PEO),33 poly(methyl methacrylate) (PMMA),34 and poly(vinylidene fluoride) (PVDF).35 PVDF, a semi-crystalline polymer with radially crystallized, ball-like structures, has chain segments that create a dipole moment due to the presence of electro-negative fluorine and electro-positive hydrogen atoms.36 This dipole moment grants PVDF a moderate dielectric constant, facilitating the dissociation of lithium salts in the electrolyte. PVDF has various crystalline phases, including α and β phases; α-phase PVDF provides thermodynamic stability, while β-phase offers a higher dielectric constant. PVDF-based SSEs are typically prepared by dissolving PVDF in a solvent, which is then dried and evaporated to form a matrix containing lithium salt.35 Research by Nan et al. has shown that the solvent molecules in the PVDF-based SSE coordinate with lithium ions, facilitating lithium-ion transport by interacting with fluorine atoms on the PVDF chains.37,38 This ability to promote ionic conductivity has made PVDF-based SSEs attractive for use with lithium metal anodes.
This review will summarize recent advances in PVDF-based flexible SSEs for lithium metal batteries, focusing on modification strategies, electrochemical performance, and design structures. We will also discuss the preparation and modification methods, the nano- and microstructures, and the electrochemical properties of PVDF-based SSEs. Finally, we will provide an outlook on the potential of PVDF-based SSEs for high-performance lithium metal batteries.
By leveraging these intrinsic PVDF properties, researchers have developed various modification techniques to engineer polymer electrolytes with enhanced and consistent performance. Key strategies for optimizing PVDF-based SSEs in LMBs include:
Doping with inorganic fillers: Incorporating inorganic fillers such as Li3PO4, Al2O3, or LLZTO improves ionic conductivity, mechanical strength, and stability. These fillers help reduce crystallinity in PVDF, creating more amorphous regions that facilitate lithium-ion mobility and improve electrochemical performance.
Blending with organic fillers: Organic additives like poly(ethylene oxide) (PEO) or poly(methyl methacrylate) (PMMA) can be blended with PVDF to increase flexibility, reduce brittleness, and improve interface compatibility with lithium metal anodes. Blending polymers also adjusts the mechanical properties, enhancing stability during battery cycling.
Inorganic/organic composite fillers: A combined approach using both inorganic and organic fillers can yield synergistic benefits, where inorganic components contribute to thermal stability and ionic conductivity, while organic additives enhance flexibility and interfacial contact with electrodes.
Chemical modification of PVDF: Chemical modifications such as grafting and functionalizing PVDF with ion-conducting groups can increase its ionic conductivity by introducing additional lithium-ion conduction pathways and lowering the glass transition temperature.
These modification strategies, outlined in Scheme 1, facilitate the design of PVDF-based SSEs that meet the specific requirements of high-performance LMBs. Table 1 summarizes the synthesis methods, compositions, and electrochemical properties of various PVDF-based SSEs tailored for LMB applications. By systematically exploring and combining these approaches, PVDF-based electrolytes can achieve the high stability, flexibility, and ionic conductivity necessary for advanced LMB technologies.
Scheme 1 Design strategies and properties of PVDF-based flexible electrolytes for high-performance all solid-state Lithium Metal Batteries (LMBs). |
Electrolyte | Synthetic methods | σa (mS cm−1) | tLi+b | LSVc (V) | MSd (MPa) | Lifespan Te h [F1f (mA cm−2), F2f (mA h cm−2)] | Battery configuration | Performance | Ref. |
---|---|---|---|---|---|---|---|---|---|
a Conductivity (room temperature).b Ion transference number.c Linear sweep voltammetry testing electrochemical stability.d Mechanical strength.e Time (h).f F1: current density (mA cm−2); F2: specific area capacity (mA h cm−2). | |||||||||
Inorganic filler | |||||||||
Nanoparticles | |||||||||
PVDF/LiClO4/DMF/Li6.75La3Zr1.75Ta0.25O12 | Solution-casting method | 0.5 | — | — | 5.92 | 160 (0.05, 0.025) | LiCoO2/Li | 150, 98% (120 cycles, 0.4C) | 39 |
PVDF/LiTFSI/DMF/Li1.3Al0.3Ti1.7(PO4)3 | Solution-casting method | 0.244 | 0.52 | 4.8 | — | 3000 (0.1) | LiNi0.6Co0.2Mn0.2O2/Li | 125, 80% (400 cycles, 0.5C) | 40 |
PVDF/LiFSI/DMF/SiO2 | Solution-casting method | 0.481 | 0.59 | — | 64 | 11000 (0.1, 0.1) | LiNi0.8Mn0.1Co0.1O2/Li | 173, (300 cycles, 0.5C) | 41 |
PVDF/LiFSI/DMF NaNbO3 | Solution-casting method | 0.556 | 0.49 | 4.7 | — | 2800 (0.1, 0.1) | LiNi0.8Mn0.1Co0.1O2/Li | 177, 67.7% (1500 cycles, 1C) | 42 |
PVDF/LiTFSI/DMF/BiFeO3 | Casting and scraping method | 0.139 | 0.35 | 4.7 | — | 2500 (0.1, 0.1) | LiNi0.8Mn0.1Co0.1O2/Li | 141.7, 89% (400 cycles, 0.3C) | 43 |
PVDF/LiTFSI/NMP/LiSnZr(PO4)3 | Solution-casting method | 0.0576 | 0.73 | 4.73 | — | 2000 (0.04) | Li4Ti5O15/Li | 133, 88% (20, 0.1C) | 44 |
PVDF/LiTFSI/DMF/Li3Zr2Si2PO12 | Solution-casting method | 0.228 | — | 5.2 | — | 600 (0.1) | LiNi0.5Co0.2Mn0.3O2/Li | 150, 88.5% (100 cycles, 0.1C) | 45 |
PVDF/LiFSI/DMF/Li6.5La3Zr1.5Ta0.1Nb0.4O12 | Solution-casting method | 0.105 | 0.66 | 4.8 | — | 400 (0.1, 0.1) | LiNi0.8Mn0.1Co0.1O2/Li | 137, (80 cycles, 0.3C) | 46 |
PVDF/LiTFSI/DMF/LiNO3 | Solution-casting method | 0.129 | 0.32 | — | — | 400 (0.2, 0.2) | LiNi0.8Mn0.1Co0.1O2/Li | 118.5, 93% (200 cycles, 0.5C) | 47 |
PVDF/LiFSI/DMF/MS zeolite | Solution-casting method | 0.45 | 0.47 | 4.6 | — | 5100 (0.1, 0.1) | LiNi0.8Mn0.1Co0.1O2/Li | 178.5, 92.7% (500 cycles, 1C) | 48 |
PVDF/LiClO4/DMF/(Mg,Al)2Si4O10(OH) | Casting and scraping method | 0.12 | 0.54 | — | 4.7 | 100 (0.05, 0.05) | Ni1/3Mn1/3CO1/3O2/Li | 117.6, 97% (200 cycles, 0.3C) | 49 |
Nanowire | |||||||||
PVDF/LiTFSI/NMP/Li0.35La0.55TiO3 | Solution-casting method | 0.53 | — | 5.1 | 9.5 | 300 (0.2, 0.2) | LiFePO4/Li | 121, 99% (100 cycles, 1C) | 50 |
PVDF/LiFSI/DMF-BaTiO3/Li0.33La0.56TiO3−x | Solution-casting method | 0.82 | 0.57 | — | 2.11 | 1900 (0.1, 0.1) | LiNi0.8Mn0.1Co0.1O2/Li | 180, 70% (1000 cycles, 1C) | 51 |
PVDF/LiClO4/DMF/V2O5 | Solution-casting method | 2.2 | 0.58 | 5.2 | — | 2500 (0.5, 0.5) | LiFePO4/Li | 150, 95.3% (300 cycles, 0.1C) | 52 |
PVDF/LiTFSI/DMF/d-HNTs | Casting and scraping method | 0.29 | 0.75 | — | 49 | 400 (0.5, 0.5) | LiFePO4/LiLiNi0.8Mn0.1Co0.1O2/Li | 148, 80% (300 cycles, 1C) | 53 |
149, 70% (200 cycles, 1C) | |||||||||
PVDF/LiTFSI/DMF/LiF/LLAZO | Casting and scraping method | 0.44 | 0.33 | 4.9 | — | 300 (0.1, 0.1) | LiFePO4/Li | 167, 96% (250 cycles, 0.5C) | 54 |
Nanosheet | |||||||||
PVDF/LiFSI/DMF/g-C3N4 | Solution-casting method | 0.69 | 0.49 | 4.7 | 11.2 | 2200 (0.1, 0.1) | LiNi0.8Mn0.1Co0.1O2/Li | 146.5, 76.6% (1700 cycles, 1C) | 55 |
PVDF/LiODFB/DMF/PC | Casting and scraping method | 0.118 | — | 4.75 | 4.3 | 160 (0.1, 0.1) | LiCoO2/Li | 125, 84% (300cycles, 0.1C) | 56 |
PVDF/LiTFSI/NMP/ISMN | Solution-casting method | 0.44 | 0.5 | 4.92 | — | 5000 (0.2, 0.1) | LiFePO4/Li | 154, 88.9% (500 cycles, 0.5C) | 57 |
PVDF/LiTFSI/NMP/h-BN | Solution-casting method | 0.29 | 0.62 | 5.24 | 3.45 | 1200 (0.1, 0.1) | LiFePO4/Li | 121.4, 96% (160 cycles, 0.2C) | 58 |
Organic filler | |||||||||
PVDF/LiFSI/DMF/PAA | Solution-casting method | 0.09 | — | 4.64 | — | 900 (0.44, 0.22) | LiCoO2/Li | 125, 97% (1000 cycles, 0.1C) | 59 |
PVDF/LiTFSI/DMSO/AMPS | Solution-casting method | 0.22 | 0.49 | 4.7 | — | 2100 (0.1, 0.1) | LiFePO4/Li | 127.6, 90.8% (220 cycles, 1C) | 60 |
PVDF/LiTFSI/DMF/TFBQ | Solution-casting method | 0.239 | 0.42 | 5.0 | — | 2000 (0.1, 0.1) | LiNi0.6Co0.2Mn0.2O2/Li | 150, 80% (180 cycles, 0.2C) | 61 |
PVDF/LiTFSI/DMF/MgPFPAA | Casting and scraping method | 0.14 | 0.34 | 4.8 | — | 2400 (0.2, 0.2) | LiFePO4/Li | 120, 74.9% (1500 cycles, 5C) | 62 |
PVDF/LiTFSI/DMF/HFA | Casting and scraping method | 0.241 | — | 4.9 | — | 1700 (0.1, 0.1) | LiNi0.6Co0.2Mn0.2O2/Li | 176.8, 80% (600 cycles, 0.2C) | 63 |
PVDF/LiFSI/DMF/FEC/LIDFP | Solution-casting method | 0.479 | 0.43 | 4.6 | 6.5 | 3000 (0.1, 0.1) | LiFePO4/Li | 148, 84% (400 cycles, 1C) | 64 |
Organic/inorganic filler | |||||||||
PVDF/LiFSI/DMF-h-PAN@MOF | Solution-casting method | 1.03 | — | 4.45 | 20.8 | 3200 (0.1, 0.1) | LiNi0.8Mn0.1Co0.1O2/Li | 150, 61.9% (1400 cycles, 2C) | 65 |
PVDF/PVAC/LiTFSI/LiBOB/DMF/Li6.4La3Zr1.4Ta0.6O12 | Casting and scraping method | 0.496 | 0.57 | 5.4 | 6 | 1300 (0.1, 0.1) | LiFePO4/Li | 145, 92.3% (220 cycles, 0.5C) | 66 |
PVDF/LiTFSI/NMP/ZIF-90 | Casting and scraping method | 0.62 | 0.48 | — | 2.1 | 1000 (0.05, 0.05) | LiFePO4/Li | 120, 95% (300 cycles, 1C) | 67 |
Other strategy | |||||||||
PVDF-OH/LiTFSI/DMSO | Solution-casting method | 0.71 | — | — | 16.1 | 1000 (0.1, 0.1) | LiFePO4/Li | 145.9, 85.4% (1000 cycles, 0.5C) | 68 |
P(VDF-CHF3-CH2FCl)/LiFSI/DMF | Solution-casting method | 0.78 | 0.57 | 4.4 | — | 11000 (0.05, 0.05) | LiNi0.8Mn0.1Co0.1O2/Li | 154, 94.9% (300 cycles, 1C) | 69 |
PVDF/LiTFSI/NMP/Li6.1Al0.3La3Zr2O12 framework | Immersing method | 0.437 | 0.72 | 5.08 | — | 1000 (0.1, 0.1) | LiNi0.6Co0.2Mn0.2O2/Li | 160, 90% (200 cycles, 0.2C) | 70 |
PVDF741/LiClO4/DMF | Casting and scraping method | 0.12 | — | 4.9 | — | 1000 (0.1, 0.1) | LiFePO4/Li | 100, 80% (500 cycles, 0.5C) | 71 |
P(VDF-TrFE-CTFE)/LiTFSI/DMF | Solution-casting method | 0.31 | 0.33 | 4.6 | — | 1200 (0.05, 0.05) | LiFePO4/Li | 146, 98.5% (150 cycles, 0.5C) | 72 |
For instance, Nan et al. developed composite solid electrolytes (CSEs) with garnet-type Li6.75La3Zr1.75Ta0.25O12 (LLZTO) nanoparticles as active fillers.39 The La atoms in LLZTO interact with nitrogen atoms and CO groups from the solvent N,N-dimethylformamide (DMF), creating electron-enriched nitrogen atoms that act as Lewis bases, leading to the dehydrofluorination of PVDF. This interaction between PVDF, lithium salt, and LLZTO enhances the flexible electrolyte's performance, resulting in a high ionic conductivity of approximately 0.5 mS cm−1 at 25 °C. In a similar study, Liu et al. fabricated flexible CSEs using a PVDF matrix with lithium salt (LiTFSI), DMF, and NASICON-type Li1.3Al0.3Ti1.7(PO4)3 (LATP) ceramic nanoparticles as fillers.40 DMF facilitates lithium salt dissociation and forms lithium-rich complexes [Li(DMF)nTFSI] with Li+, which exhibit ionic liquid-like characteristics that improve conductivity. This design yielded a LATP-PVDF/Li CSE with a high ionic conductivity of 0.244 mS cm−1 and an electrochemical stability window of up to 4.8 V (vs. Li+/Li).
However, the thickness and limited mechanical strength of PVDF-based SSEs can restrict their practical applications. To address this, Ma et al. developed an ultrathin PVDF-based SSE by integrating a 7 μm polyethylene (PE) separator and SiO2 nanoparticles with silicon hydroxyl (Si-OH) groups.41 This composite, known as PPSE, achieved a total thickness of only 20 μm and exhibited ultra-high mechanical strength (64 MPa). The nano-SiO2 particles anchored DMF molecules, enhancing ion conductivity in PVDF and preventing side reactions with lithium metal. This modification increased the ionic conductivity to 0.48 mS cm−1, reduced activation energy (0.19 eV), and achieved a high lithium transference number (0.59).
Introducing functional fillers with ferroelectric or dielectric properties has also proven effective in improving PVDF-based SSEs. For instance, Kang et al. utilized dielectric NaNbO3 nanoparticles in a PVDF-based SSE (PNNO-5), which induced the formation of a high-dielectric β-phase in PVDF (Fig. 1a).42 This phase enhances Li+ coordination with FSI− anions due to polarized F atoms, which aid in abundant and mobile Li+ ion formation (Fig. 1b–d). The PNNO-5 SSE achieved a high ionic conductivity of 0.556 mS cm−1, with a decreased ion migration activation energy of 0.22 eV, as compared to 0.33 eV in unmodified PVDF SSEs (Fig. 1e). This configuration allowed a LiNi0.8Mn0.1Co0.1O2 (NCM811)/PNNO-5/Li cell to retain 67.7% capacity over 1500 cycles at 1C and demonstrated a high initial capacity of 177.2 mA h g−1 with 95% retention over 100 cycles at 0.5C (Fig. 1f).
Fig. 1 Introducing functional fillers with ferroelectric or dielectric properties to improve PVDF-based SSEs. (a) Surface SEM images and optical photographs of PNNO-5. (b and c) Raman spectra of PVDF and PNNO-5. (d) Schematic illustration showing Li salt dissociation facilitated by the β-phase of PVDF. (e) Arrhenius plots comparing ionic conductivity of PVDF and PNNO-5 electrolytes. (f) Cycling performance of NCM811/PNNO-5/Li pouch cell. Reproduced with permission.42 Copyright 2024, Advanced Materials. (g) Schematic illustration of the dipole effect on the conduction mechanism. (h) Cycling performance of Li/NCM811 batteries. Reproduced with permission.43 Copyright 2024, Journal of Materials Chemistry A. |
Li et al. also introduced ferroelectric BiFeO3 nanoparticles to modify PVDF-based SSEs, which helped distribute the electric field more uniformly at the electrolyte/electrode interface, resulting in uniform Li deposition.43 The transference number of Li+ increased from 0.18 to 0.35 compared to pure PVDF, enhancing overall battery performance. The BiFeO3 particles’ spontaneous polarization along the [111] direction generates dipoles that interact electrostatically with TFSI− anions (Fig. 1g), further dissociating LiTFSI and reducing the migration resistance of [Li(DMF)x]+ complexes. The assembled Li/NCM811 full cells showed an excellent performance, which a high coulombic efficiency of 99% and a capacity retention of 89% achieved after 400 cycles (Fig. 1h).
Compared to granular fillers, ceramic nanowires can substantially enhance the conductivity of SSEs by creating continuous pathways that facilitate lithium-ion transport. The interconnected nanowire network also improves the electrolyte's mechanical strength, thus enhancing battery stability and safety.73 For instance, Su and colleagues developed a CSE with Li0.35La0.55TiO3 nanofibers and PVDF/LiTFSI, achieving an ionic conductivity of 0.53 mS cm−1 at room temperature and a voltage window up to 5.1 V.50 Shi et al. synthesized BaTiO3-Li0.33La0.56TiO3−x (BTO-LLTO) nanowires with a heterojunction structure through electrospinning, followed by calcination at 1000 °C.74 Introducing these nanowires into a PVDF matrix yielded a PVDF-based CSE (PVBL) with high ionic conductivity (0.82 mS cm−1) and lithium transference number (0.57) (Fig. 2a–c). The PVBL exhibited stable cycling for over 1900 hours at 0.1 mA cm−2 and 0.1 mA h cm−2 in a symmetrical cell (Fig. 2d) and achieved a capacity retention of 70% after 1000 cycles in an NCM811/PVBL/Li battery (Fig. 2e).
Fig. 2 Incorporating an interconnected nanowire network to enhance PVDF-based SSEs. (a) Illustration of the Li salt dissociation and Li+ transport facilitated by the coupled BTO-LLTO in the PVBL electrolyte. (b) Schematic of Li salt dissociation by a polarized dielectric material. (c) The ssNMR spectra of the PVL and (d) PVBL electrolytes before and after cycling in a 6Li symmetric cell, with (e) corresponding peak ratios of 6Li+ transport paths. (f) Ionic conductivities of the PVDF-based electrolytes at 25 °C. (g) Long-term cycling performance of PVBL in symmetric cells. Reproduced with permission.51 Copyright 2017, Nature Nanotechnology, Nature Publishing Group. |
In another study, He et al. incorporated g-C3N4 nanosheets (GCNs) into a PVDF-based SSE, achieving high ionic conductivity of 0.69 mS cm−1.75 During cycling, GCNs react with lithium metal to form a Li3N-enriched SEI layer, significantly reducing side reactions and ensuring rapid charge transfer. The GCNs’ adsorption capacity for residual DMF further improves electrochemical stability, enabling the Li symmetrical cell to cycle steadily for 2200 hours at 0.1 mA cm−2 and 0.1 mA h cm−2. Additionally, the NCM811 cathode-based LMBs exhibited a high discharge capacity of 108 mA h g−1 at 5C and excellent cycling stability over 1700 cycles at 1C.
In summary, doping PVDF-based SSEs with various inorganic fillers – ranging from 0D nanoparticles to 1D nanowires and 2D nanosheets – enhances their electrochemical properties by increasing ionic conductivity, mechanical strength, and stability, facilitating promising applications in high-performance lithium metal batteries.
To further stabilize the interface, Zhang et al. incorporated isosorbide mononitrate (ISMN), a functional additive with a non-resonant structure (O2–N–O–), into PVDF-based SSEs to create a stable N-rich solid electrolyte interface, reducing Li dendrite formation and side reactions (Fig. 3a).57 ISMN, when added to PVDF-based SSEs, forms a stable interface upon lithium stripping/plating by cleaving the N–O bond in its non-resonant structure, yielding a nitrogen-rich layer that enhances ultra-stable flexible SSLMBs. The Li/Li symmetric cell achieved stable Li stripping/plating cycling for over 5000 hours at a current density of 0.2 mA cm−2 and capacity of 1.0 mA h cm−2. Additionally, the Li|LiFePO4 cell exhibited an initial discharge capacity of 154.0 mA h g−1, with a capacity retention of 88.9% after 500 cycles at 0.5C. The flexible NCM622/Li pouch cell maintained a high discharge capacity retention of 97.2% over 100 cycles at 0.5C.
Fig. 3 Enhancing PVDF-based SSE through blending with organic fillers. (a) Schematic illustration of the modification mechanism using ISMN for PVDF-based SSE. Reproduced with permission.57 Copyright 2024, Advanced Functional Materials. (b) Schematic illustration showing the role of AMPS in promoting Li+ conduction and suppressing Li dendrite formation. (c) FTIR spectra of AP-PE. (d) SEM images of the AP-PE membrane. (e) SEM images of Li electrodes from Li/AP-PE after 500 hours of cycling at 0.1 mA cm−2 and 0.1 mA h cm−2. (f) Cycling performance of Li/AP-PE/LFP batteries at 0.5C and (g) the corresponding charge–discharge voltage profiles. Reproduced with permission.60 Royal Society of Chemistry 2022. |
Moreover, Nan et al. developed a PVDF-based SSE with 2-acrylamido-2-methylpropane sulfonic acid (AMPS) as an additive to enhance Li+ conduction by reducing the crystallinity of PVDF and immobilizing anions.60 AMPS induces the formation of a LiF/Li2Sx/Li2SO3/Li3N-rich interface at the Li anode, suppressing dendrite growth (Fig. 3b–d). The resulting AMPS-PVDF polymer electrolyte (AP-PE) enables symmetric cells to cycle stably for 2100 hours at 0.1 mA cm−2 and 0.1 mA h cm−2. After 500 hours of cycling, the Li anode in these symmetric cells exhibited a clean surface with no dendrite formation (Fig. 3e). The Li/AP-PE/LFP cell demonstrated stable cycling performance, retaining 90.8% of its capacity after 200 cycles at 0.5C (Fig. 3e–g).
By blending PVDF-based SSEs with various organic fillers, researchers have significantly enhanced ionic conductivity, interface stability, and cycling performance, making these electrolytes viable candidates for high-performance, long-lasting LMBs.
Fig. 4 Inorganic/organic composite fillers for enhanced PVDF-based SSE. (a) Schematic of ion transport and solvation structures in PVDF and PPM electrolytes. (b) SEM images of h-PAN@MOF networks. (c) 7Li ss-NMR spectra. (d) 19F ss-NMR spectra. (e) Galvanostatic cycling curves of Li||Li symmetric cells with PVDF, PP, and PPM electrolytes. (f) Long-term cycling stability of Li||NCM811 cells. Reproduced with permission.51 Copyright 2023, Nature Nanotechnology, Nature Publishing Group. |
The h-PAN@MOF network also imparts high mechanical strength (20.84 MPa) to the electrolyte, effectively inhibiting lithium dendrite growth. The adsorption energy of the CO group in DMF on the MOF crystal (−1.11 eV) is three times greater than that on the PVDF chains (−0.35 eV). This strong adsorptive interaction between DMF and MOF realigns solvent molecules around the h-PAN@MOF networks, significantly reducing DMF decomposition at the PPM electrolyte/lithium metal interface.
Solid-state nuclear magnetic resonance (ss-NMR) spectroscopy of the 7Li and 19F spectra was conducted to examine the interactions between FSI-Li+ and DMF-Li+ complexes. An upfield shift in the 7Li spectrum indicates an enhanced shielding effect of lithium nuclei in the PPM electrolyte, suggesting tighter coordination with surrounding ligands (Fig. 4c). Changes in chemical shifts in the 19F spectra of FSI also indicate a weakened DMF-Li+ interaction and stronger FSI-Li+ coordination (Fig. 4d). The combined effects of h-PAN and MOF establish a competitive Li+ coordination environment and alter solvation structures to promote rapid, linear Li+ transport. This synergy enhances interfacial stability with electrodes.
As a result, the Li|PPM|Li battery demonstrated stable voltage hysteresis over 3200 hours at a current density of 0.1 mA cm−2 and a capacity of 0.1 mA h cm−2 (Fig. 4e). Additionally, the Li|PPM|NCM811 battery retained 61.9% of its capacity over 1400 cycles at 2C (Fig. 4f). This promising performance highlights the effectiveness of inorganic/organic composite fillers in enhancing the stability and ionic conductivity of PVDF-based SSEs in lithium metal batteries.
In another advancement, Huang et al. demonstrated that Li+ can transport through the crystalline phase of PVDF by incorporating dipolar defects into the crystals.69 By adding trifluoroethylene (CHF3) and chlorofluoroethylene (CH2FCl) as dipolar defects into VDF crystals, they triggered rapid Li+ movement through ion–dipole interactions. These defects expanded the interchain distance in PVDF from 4.39 to 4.83 Å, facilitating –CH2CF2 dipole vibrations at room temperature and supporting Li+ migration through ion-dipole interactions, thereby transforming PVDF into a fast ion conductor (Fig. 5a). As a result, the defective PVDF (d-PVDF) SSE achieved a high ionic conductivity of 0.78 mS cm−1 and a lithium-ion transference number (tLi+) of 0.57 (Fig. 5d).
Fig. 5 Chemical modification of PVDF for enhanced PVDF-based SSE. (a) Schematic illustrating Li+ transport within d-PVDF crystals. (b) Bipolar D–E loops of PVDF and d-PVDF. (c) Schematic representation of Li+ and FSI− location within d-PVDF SPE. (d) Lithium-ion transference number (tLi+) comparison between PVDF and d-PVDF SPE. (e) Cycling performance of Li//Li symmetrical cells with d-PVDF and PVDF SSE. Reproduced with permission.69 Copyright 2024, Energy & Environmental Science. |
Further analysis with displacement-electric field (D–E) measurements showed that d-PVDF, unlike ferroelectric (FE) PVDF, exhibited a slim hysteresis loop, indicating a significant reduction in coercive field from 67.3 to 10.7 MV m−1 (Fig. 5b). This observation confirmed the presence of –CHF3 and –CH2FCl dipolar defects in the VDF crystals, which promote Li+ integration within the crystals, transforming defective d-PVDF crystals into single-ion conductors (Fig. 5c). The Li/d-PVDF/Li symmetrical cell demonstrated an extended lifespan of 11000 hours (450 days) at 0.05 mA cm−2 (Fig. 5e). Furthermore, NCM811/d-PVDF/Li batteries exhibited excellent capacity retention of 94.9% after 300 cycles at 1C.
These findings underscore the potential of modifying PVDF to create fast ion-conducting pathways, achieving significant improvements in ionic conductivity, lithium-ion transference, and overall stability in PVDF-based SSEs for advanced lithium metal batteries.
Fig. 6 Additional strategies for enhancing the performance of PVDF-based SSE. (a) Schematic illustration of ceramic-based CSE. (b) Cross-sectional SEM image and magnified view of the selected region of the CSE. (c) Illustration of the interactions between PVDF-LiTFSI CSE and porous LLZO framework. (d) Charge–discharge voltage profiles of a Li/ceramic-based CSE/NCM622 solid-state battery. (e) Cross-sectional SEM and corresponding EDS mappings of the SSB before and after cycling at 3–4.8 V. Reproduced with permission.70 Copyright 2024, Cell Reports Physical Science. |
The ceramic skeleton creates a 3D conductive network with micron-sized pores that allow solubilized SSEs to permeate (Fig. 6b). The resulting SSE has high porosity (45.74%) and a wide voltage window (5.08 V), making it compatible with high-voltage cathodes. The ceramic-based SSE, with a high ceramic-mass composition (93%), provides robust mechanical support that suppresses lithium dendrite formation, while the porous LLZO framework allows for substantial PVDF-LiTFSI loading (Fig. 6c).
When used in LiNCM622/ceramic-based SSE/Li full solid-state lithium metal batteries (SSLMBs), the CSE demonstrated stable cycling over 200 times within a voltage range of 3 to 4.8 V, maintaining a high coulombic efficiency of 99.76% at 0.2C (Fig. 6d). After 200 cycles, the interfacial zone between the cathode and CSE remained intact, with no delamination or cracking. Additionally, an F− and N-rich interface formed between the cathode and ceramic-based CSE, enhancing stability and performance (Fig. 6e). This 3D scaffold approach highlights the potential for creating stable, high-performance PVDF-based SSEs suitable for high-voltage and long-cycle SSLMB applications.
In conclusion, this review provides a foundational understanding and recent advancements in the design and use of PVDF-based SSEs for SSLMBs. It will inform future efforts in developing high-performance PVDF-based SSLMBs and help accelerate their practical applications.
This journal is © The Royal Society of Chemistry 2025 |