Polymer Electrolytes for High Energy Density Ternary Cathode Material-Based Lithium Batteries
- 91 Downloads
Abstract
Layered transition metal oxides such as LiNixMnyCo1−x−yO2 and LiNixCoyAl1−x−yO2 (NCA) (referred to as ternary cathode material, TCM) are widely recognized to be promising candidates for lithium batteries (LBs) due to superior reversible capacities, high operating voltages and low production costs. However, despite recent progress toward practical application, commercial TCM-based lithium ion batteries (LIBs) suffer from severe issues such as the use of flammable and hazardous electrolytes, with one high profile example being the ignition of NCA-based LIBs used in Tesla Model S vehicles after accidents, which jeopardizes the future development of TCM-based LBs. Here, the need for TCM and flammable liquid electrolytes in TCM-based LBs is a major obstacle that needs to be overcome, in which conflicting requirements for energy density and safety in practical application need to be resolved. To address this, polymer electrolytes have been demonstrated to be a promising solution and thus far, many polymer electrolytes have been developed for high-performance TCM-based LBs. However, comprehensive performances, especially long-term cycling capabilities, are still insufficient to meet market demands for electric vehicles, and moreover, comprehensive reviews into polymer electrolytes for TCM-based LBs are rare. Therefore, this review will comprehensively summarize the ideal requirements, intrinsic advantages and research progress of polymer electrolytes for TCM-based LBs. In addition, perspectives and challenges of polymer electrolytes for advanced TCM-based LBs are provided to guide the development of TCM-based power batteries.
Graphical Abstract
Keywords
Lithium batteries Ternary cathode material All-solid-state polymer electrolyte Gel polymer electrolyte1 Introduction
Approximate range of average discharge potentials and specific capacities of common intercalation-type cathodes obtained through experimental results. The acronyms for the cathode active materials are: lithium cobalt oxide (LCO), lithium manganese oxide (LMO), lithium cobalt phosphate (LCP), lithium iron phosphate (LFP), lithium iron fluorosulfate (LFSF), and lithium titanium sulfide (LTS) [33].
Copyright 2015, Elsevier
- (1)Cation mixing: TCM mainly exhibits O1 and O3 crystal structures as well as a hybrid crystal phase of the two which is referred to as H1-3, which forms during partial lithium deintercalation (Fig. 2a–c) [34, 35, 36]. Here, all crystal textures show a close and well-organized oxygen pattern along with interstitial sites among alternating layers occupied by lithium ions and transition metal (M = Ni, Co, Mn and Al) cations, in which a formula of LixMO2 can be used to represent the crystal structures of the TCM. However, Ni2+ ions can easily occupy 3b lithium sites due to a similar ionic radius to Li+ [37], and other transition metal ions such as Mn2+ can also appear in the cation mixing layer. This cation mixing can occur during the synthesis process of TCM as well as electrochemical operations and is detrimental to the electrochemical performances of TCM-based LBs.Fig. 2
Related crystal structures for TCM. Blue octahedra represent MO6 units and green octahedra/tetrahedra represent Li sites [22].
Copyright 2017, Wiley
- (2)
Phase transformation: O3 crystal frameworks in TCM can irreversibly transform from O3 to spinel or disordered rock-salt motifs during operation (Fig. 2d–e). This transformation is thermodynamically favorable during the delithiation process [37] and is caused by the migration of metal ions (~ 25%) from the M cation layer to the particular octahedral sites in the lithium ion layer as well as the subsequent densification of the layered MO2 into a rock-salt MO along with the release of oxygen [33, 34]. Moreover, these irreversible phase transformations can generate thin coatings onto TCM surfaces and kinetically limit Li diffusion across the reconstruction phase, causing detrimental impacts on the long-term cycling and C-rate ability of TCM-based LBs [22].
- (3)
Trade-offs between high energy density and high stability: Varying electrochemical configurations of TCM based on varying ratios of transitional metal cations tend to exhibit varying electrochemical and thermal properties. For example, the increasing ratios of Ni cations in NMCs can enhance working voltages and specific capacities but lead to inferior cycling stabilities and rate performances as a result of severe Ni/Li cation mixing. In addition, Ni cation content is inversely proportional to thermal stability in NCMs [38, 39, 40]. In another example, increasing Mn cation content leads to increased operational safety but at the expense of the formation of more spinel phases. As for Co, although Co cations can effectively mitigate disordered crystal textures resulting from Ni/Li cation mixing and gain good cycling and rate performances, higher Co content leads to decreased cell capacity [23]. Furthermore, in the comparison between NMC and NCA, NCA can provide higher rate capabilities than NMC due to high electron and ion-conductive characteristics; and similar to Mn cations in NMC, Al cations in NCA can demonstrate inactive electrochemical properties and enable NCA to possess better cycling stabilities. However, despite the presence of synergistic effects of transitional metal cations in TCM, both NMC and NCA deliver relatively inferior cyclabilities due to irreversible side reactions and mechanical degradation on the electrode surface at high working potentials and elevated temperatures. In addition, higher energy density TCM tends to exhibit inferior structure stability and electrochemical performance.
- (1)Safety hazards: Uncontrollable side and exothermic reactions between liquid electrolytes and oxygen species released from structurally unstable cathode active materials during cycling and especially at elevated temperatures or high voltages [26] in TCM-based LIBs can cause extreme reactions as seen in accidents involving Tesla electric vehicles. In addition, volatile and flammable liquid electrolytes can greatly contribute to thermal runaway, causing small ignitions to grow significantly, in which chain reactions during thermal runaway of LIBs have been extensively studied from qualitative interpretations (Fig. 3) [41]. And as a result, safety hazards caused by traditional organic electrolytes are a major challenge in the commercialization of TCM-based LIBs.Fig. 3
Qualitative interpretation of the chain reactions during thermal runaway [41].
Copyright 2018, Elsevier
- (2)
Parasitic side reactions between cathodes and electrolytes: The parasitic and irreversible side reactions between cathodes and liquid electrolytes, as well as the continuous oxidative decomposition of liquid electrolytes on cathode surfaces are intrinsic flaws in which the dissolution of transition metal ions from TCM is thought to be one of the dominant factors and is responsible for the poor cycling lifespan and capacity decay in TCM-based LIBs. Here, acidic species from liquid electrolytes such as PF5 and HF generated from side reactions of LiPF6 can severely corrode TCM and facilitate the significant dissolution of transition metal cations, especially at elevated temperatures or high operating voltages [42, 43, 44, 45], increasing the impedance of the cathode and electrolyte interface with the accumulation of dissolution side products causing capacity decay. In addition, carbonates in liquid electrolytes are highly reactive toward oxygen species generated from cathodes and acidic compounds such as HF evolving from LiPF6 to produce complex side products including polycarbonates, semi-carbonates and so on [46, 47], impeding lithium ion transfer and deteriorating cell performances. Furthermore, transition metal ions can also act as Lewis acids to activate these side reactions, leading to severe decomposition of organic solvents, which is again accelerated under high temperatures or highly delithiated states.
- (3)
Unquenchable lithium dendrite formation: Continuous lithium dendrite formation restricts the application of liquid electrolytes in lithium metal batteries in which lithium dendrite growth in liquid electrolytes can penetrate separators and cause short circuiting and subsequent safety issues [48]. In addition, liquid electrolytes lack mechanical strength and cannot impede lithium dendrite formation.
- (4)
Severe cathode/anode interactions: The liquidity of electrolytes can increase interactions between cathodes and anodes and has a harmful effect on anode intercalation/deintercalation chemistry [45]. For example, transition metal ions (i.e., Mn2+) dissolved from the cathode can diffuse across the electrolyte and the separator and deposit onto the anode. Furthermore, the ion-exchange process between dissolved Mn species in the electrolyte and lithium ions deposited on the anode can further accelerate the erosion of cathode lattices [49, 50]. Additionally, chemical crossover of oxygen released from the cathode can cause thermal runaway without internal short circuiting [41].
Given these internal drawbacks of electrolytes, many strategies such as surface cathode modification and additive introduction have been employed to enhance the electrochemical performance and safety of traditional TCM-based LIBs [51, 52, 53, 54, 55, 56, 57, 58, 59, 60, 61, 62]. In addition, flame-retardant electrolytes are also being investigated, such as flame-retardant solvents [63, 64, 65, 66, 67, 68] and highly concentrated electrolytes [69, 70, 71, 72, 73, 74, 75, 76]. However, the use of these low flammable electrolytes in LBs again presents compromises between electrochemical performance, costs and safety. Therefore, the balance between energy density and safety is a major issue in the practical application of TCM-based batteries in which polymer electrolytes have been widely considered to be an alternative solution [77, 78, 79, 80, 81, 82, 83, 84, 85]. This is because compared with liquid electrolytes, polymer electrolytes exhibit advantages such as high safety, transition metal ion dissolution suppression from cathodes, lithium dendrite suppression close to anodes and interaction reduction between cathodes and anodes [86, 87, 88, 89]. In addition, due to the higher safety tolerance of lithium metal anodes in comparison with liquid electrolytes [90], polymer electrolytes can couple with lithium metal anodes to provide higher energy density and lower cost TCM-based LBs [80].
And as a result, many types of polymer electrolytes have been explored for high-performance TCM-based LBs. However, although polymer electrolytes possess good battery performances, comprehensive performances such as long-term cycling stability are still insufficient to meet electric vehicle demands. In addition, reviews on polymer electrolytes specifically for TCM-based LBs are lacking. Because of this, this review will present the ideal requirements and natural strengths of polymer electrolytes for TCM-based LBs and provide detailed discussions into the advantages of various polymer electrolytes (e.g., polymers containing ethoxylated (EO) motifs, polyacrylates, polycarbonates, vinylidene fluoride (VDF)-based polymers, polyacrylonitriles (PAN), as well as aromatic polymers). In addition, perspectives and potential challenges of polymer electrolytes for practical application in TCM-based LBs are presented to guide future development of polymer electrolyte-based LBs using TCM, especially for power batteries.
2 Ideal Requirements, Inherent Advantages and Recent Progress of Polymer Electrolytes for TCM-Based LBs
- (1)
High ionic conductivities of more than 10−4 S cm−1 for APEs and more than 10−3 S cm−1 for GPEs at room temperature: Ionic conductivity is the dominant parameter of electrolytes, strongly influencing electrochemical behavior, and is tied to the rate capability of battery cells.
- (2)
A relatively wide electrochemical stability window (> 4.4 V): This factor can endow polymer electrolytes with high anti-oxidation abilities if used with high-voltage TCM-based cathodes.
- (3)
Superior mechanical strength: As reported by Monroe and Newman [96], polymer electrolyte films with a shear modulus of 6 GPa can restrain lithium dendrite generation and growth. Based on this, polymer electrolytes with robust mechanical strength can significantly restrain lithium dendrite formation in LBs.
- (4)
Tight interfacial contact and good compatibility between electrodes and electrolyte systems: This factor can enhance cycling stabilities and provide high capacity retention during operation.
- (5)
Thermal stability and chemical inertness of cell components during operation: Thermal stability enables operation at wide temperature ranges, and chemical inertness ensures high electrochemical performance.
- (6)
A high lithium ion transference number (t Li + ) close to 1: High t Li + can lead to superior power densities by suppressing the concentration polarization of polymer electrolytes.
Schematic illustration of the inherent advantages of a polymer electrolyte-based LBs using TCM as the cathode active material as compared with b corresponding LBs using liquid electrolytes
- (1)
High safety: The superior safety characteristic of polymer electrolytes can be ascribed to the lack of electrolyte leakage which is present with liquid electrolytes, and is an inherent characteristic.
- (2)
Suppression of transition metal ion dissolution from cathodes: Polymer electrolytes can act as robust barriers between TCM cathodes and liquid electrolytes and can significantly suppress interfacial side and exothermic reactions including the dissolution of transition metal ions. This characteristic enhances electrochemical performances and safety properties.
- (3)
Suppression of lithium dendrite formation: Polymer electrolytes possess high mechanical strength and enhanced uniformity in the lithiation/delithiation process, suppressing dendrite formation.
- (4)
Reduction of cathode/anode interaction: Polymer electrolytes as a solid-state form can restrain chemical shuttling between electrodes, especially for transitional metal cations, which are closely tied to the Lewis base sites in the polymer framework.
Overall, the inherent advantages of polymer electrolytes indicate great potential in TCM-based LB applications as compared with liquid electrolytes. And according to the different structural characteristics of polymer matrixes, advances in polymer electrolyte-based LBs using TCM can be classified into six types: polymers containing EO motifs, polyacrylates, polycarbonates, VDF-based polymers, PAN, and aromatic polymers.
2.1 Polymers Containing Ethoxylated Motifs
Reported EO-containing polymer electrolytes containing ethoxylated motifs and their electrochemical properties
Polymer matrix | Organic solvents/lithium salt/other additives | Ionic conductivity (S cm−1) | ESW versus Li (V) | Cell configuration | Voltage region versus Li (V) | Year | Ref. |
---|---|---|---|---|---|---|---|
PEO | LiBOB/MgO | – | 4.5 | LiNi1/3Mn1/3Co1/3O2/lithium metal | 3.0–4.2 | 2011 | [97] |
SiO2–PEO | LiTFSI | – | 4.2 | LiNi0.6Mn0.2Co0.2O2/lithium metal | 3.0–4.3 | 2018 | [98] |
P(EO/MEEGE) | LiTFSI | – | – | LiNi1/3Mn1/3Co1/3O2/graphite | 2.5–4.2 | 2017 | [99] |
Wheat flour/PEO | LiTFSI | 2.62 × 10−5 at 25 °C | 4.2 | Li(Ni0.8Co0.1Mn0.1)O2/lithium metal | 2.0–4.2 | 2017 | [100] |
P(P(EO–PO)) | EC–GBL/LiBF4 | 3.2 × 10−3 at 25 °C | 4.5 | LiNi0.8Co0.15Al0.05O2/graphite | 3.0–4.3 | 2004 | [102] |
(P(POAGA)) | EC–PC–EMC–DEC/LiPF6 | 6.2 × 10−3 at 25 °C | 5.0 | LiNi1/3Mn1/3Co1/3O2/graphite | 2.8–4.4 | 2006 | [103] |
P(POAGA) | EC–PC–EMC–DEC/LiPF6 | 6.2 × 10−3 at 25 °C | – | LiNi1/3Mn1/3Co1/3O2/graphite | 2.8–4.4 | 2006 | [104] |
P(PEODMA) | BMP-TFSI-VC/LiTFSI | 1 × 10−3 at 25 °C | 4.6 | LiNi1/3Mn1/3Co1/3O2/lithium metal | 3.0–4.3 | 2013 | [105] |
P(DPHA-PEGMEM) | EC–DEC/LiPF6 | – | – | LiNi1/3Mn1/3Co1/3O2/graphite | 3.0–4.2 | 2016 | [107] |
P(EO/MEEGE) | LiBF4 | – | – | LiNi1/3Mn1/3Co1/3O2/lithium metal | 2.5–4.2 under 60 °C | 2013 | [108] |
P(EO/MEEGE) | LiBF4 | – | – | LiNi1/3Mn1/3Co1/3O2/lithium metal | 3.0–4.2 under 50 °C | 2017 | [109] |
P(EO/EH) | LiTFSI | 1.8 × 10−5 at 60 °C | – | LiNi0.5Mn0.2Co0.3O2/lithium foil | 3.0–4.3 | 2014 | [110] |
P(EO/MEEGE/AGE) | LiTFSI | 10−4 at 60 °C | – | LiNi1/3Mn1/3Co1/3O2/graphite | 4.2 | 2014 | [111] |
P(MA475–DMA550) | P14FSI/LiTFSI | 4 × 10−4 at 25 °C | 4.8 | LiNi1/3Mn1/3Co1/3O2/lithium metal | 3.0–4.2 | 2016 | [112] |
EC–EMC–FEC/LiPF6 | – | – | LiNi0.5Mn0.2Co0.3O2/lithium metal | 2.5–4.3 | 2017 | [113] |
2.1.1 Polyoxyethylene
Polyoxyethylene (PEO)-based electrolytes have been applied in TCM-based LBs due to high lithium ion-solvation abilities, low glass transition temperatures and good interfacial compatibilities. However, these electrolytes suffer from low ionic conductivities at room temperature because conductivity is governed by the segmental motion of the polymer chains in the polymeric amorphous region. Here, the addition of nanosized inorganic fillers can provide facile transfer pathways for lithium ions and reduce the glass transition temperature of pristine PEO, thus endowing these polymer electrolytes with enhanced ionic conductivities. In addition, improved electrochemical properties such as anodic stability can also be achieved with the addition of inorganic fillers. For example, Lu et al. [97] reported that the use of PEO20–LiBOB-nanosized MgO as the electrolyte produced a high anodic stability of 4.5 V and that the corresponding LiNi1/3Co1/3Mn1/3O2/Li cell delivered a superior 1st discharge capacity of 156.8 mAh g−1 and a higher capacity retention (91% after 20 cycles at 0.2 C) at 80 °C than cells without inorganic fillers. The researchers in this study also reported the important role of LiBOB in the passivated and stable cathode/electrolyte interface (CEI). Recently, Archer et al. [98] reported a solid-state hybrid polymer electrolyte (SiO2–PEO/LiTFSI) composed of silica nanoparticles covalently grafted with a PEO chain (5 kDa) and reported that this type of soft glassy solid polymer material endowed their assembled LiNi0.6Co0.2Mn0.2O2/Li cell with an initial discharge capacity of 178 mAh g−1 and a capacity retention of over 97% after 25 cycles through a simple CEI-forming process by first wetting the cathode with a LiBOB-containing liquid electrolyte (0.4 M LiBOB, 0.6 M LiTFSI, 0.05 M LiPF6 in EC/DMC). Here, the researchers also indicated that the SiO2 nanocores with a uniform multilayered arrangement can passivate the Li anodes without limiting ionic access to the electrode.
Aside from the addition of inorganic fillers, inorganic coatings sandwiched between the cathode sheet and the electrolyte can also enhance the electrochemical performance of PEO-based APE LBs. For example, Seki [99] reported that LiNi1/3Co1/3Mn1/3O2/P(EO/MEEGE)–LiTFSI/graphite cells can produce a high initial discharge capacity (128 mAh g−1) and acceptable capacity retention (78% over 20 cycles) at 30 °C by coating LiAlO2 onto the LiNi1/3Co1/3Mn1/3O2 surface, clearly outperforming the pristine cell. In another study, Lin et al. [100] also reported that the blending of PEO polymers can effectively enhance performances through synergistic effects. Here, the researchers prepared a biocompatible and biodegradable APE through the mixing of natural wheat flour with PEO and LiTFSI and reported that the resulting APE produced an ionic conductivity of 2.62 × 10−5 S cm−1 at 25 °C with a tLi+ of 0.51. In addition, the researchers reported that their synthesized cells based on high-voltage LiNi0.8Co0.1Mn0.1O2 demonstrated excellent rate performances (a capacity of 132.3 mAh g−1 at 10 C) at 100 °C, suggesting the potential application of natural materials in high-voltage energy storage systems.
Aside from low ionic conductivities at room temperature, all-solid-state polymer batteries using PEO as the polymer host often operate at higher temperatures than its melting temperature, resulting in polymer electrolytes with weak mechanical strength. In addition, narrow electrochemical stability windows restrict application in high-voltage TCM-based cells. And although these drawbacks can be mitigated through the introduction of inorganic fillers, inorganic coatings, and blending with other polymers, weaknesses caused by the oxidative decomposition of pristine PEO cannot be eradicated. Therefore, to resolve this issue, copolymerization approaches have been developed to optimize pristine PEO.
2.1.2 Polymers with a Cross-Linked Network
In general, most polymer electrolytes comprised of only polyethylene oxide (PEO) are electrochemically unstable above 4.0 V versus Li/Li+ and usually decompose around 3.6 V [101]. To address this, copolymerization with other monomers is an effective method to enhance the electrochemical stability windows and ionic conductivities of corresponding polymer electrolytes. Furthermore, GPEs with cross-linked network structures through copolymerization tend to possess desirable mechanical properties and large uptake of liquid electrolytes.
In one example, Oh and Amine [102] studied a cross-linked GPE based on a polymer (P(P(EO–PO))) containing ethylene oxide (EO) and propylene oxide (PO) motifs and reported an electrochemical stability window of up to 4.5 V, with the resulting cell exhibiting a high capacity of more than 150 mAh g−1 at 0.2 C and decent rate performances (139 mAh g−1 at 2 C). However, the researchers also reported that these batteries possessed inferior cycling stabilities; especially at 50 °C, due to resistances arising from the CEI. In another example, Kim et al. [103] developed a GPE containing a cross-linked network of polyoxyalkylene glycol acrylate (POAGA) which demonstrated enhanced electrochemical stabilities up to 5.0 V versus Li/Li+. Here, the researchers reported that the POAGA-based cells exhibited a capacity of 146.3 mAh g−1 at 2 C rate under ambient temperatures and 109.7 mAh g−1 at 0.5 C rate at − 20 °C, as well as good cycling stabilities with a retention of 89.7% of the initial capacity after the 300th cycle. In addition, overcharge tests were also conducted in this study and demonstrated that this polymer electrolyte system provided good safety as well. POAGA and its methylated derivative-based polymer electrolytes have also been investigated by other groups to match with LiNi1/3Co1/3Mn1/3O2 cathodes and have also been reported to demonstrate good cyclability and safety properties [104, 105].
Despite these advantages, however, inferior cycling capabilities, especially at elevated temperatures, which predominately result from cathode degradation, significantly hinder application in electric vehicles [106]. To solve this issue, Jung et al. [107] developed a GPE (P(DPHA-PEGMEM)) through the copolymerization of dipentaerythritol hexaacrylate (DPHA) with poly (ethylene glycol) methyl ether methacrylate (PEGMEM) and reported that assembled batteries based on this GPE demonstrated superior cyclability at 80 °C (89% capacity retention at 3 C after 50 cycles). Here, the researchers attributed the rate performance partly to the buffering effect of lithium ion dissolution being extracted from active materials under high rate operations by polymer electrolytes. Furthermore, the researchers also reported that the ionic conductivity of their GPE was less susceptible to temperature changes and possessed enhanced thermal stabilities (up to 210 °C).
Schematic diagram of the double surface modification on both electrodes and cycling performance of a properly designed cell. Operation conditions: 4.2/2.5 V, (48 cycles (C/4)/2 cycles (C/8)) × n loops. Operation temperature: 50 °C [109].
Copyright 2017, Elsevier
In addition to the surface coating of cathodes, compositing cathodes with ion-conductive materials can also enhance the electrochemical capability of polymeric cells, and especially rate ability. For example, Nan et al. [110] studied a composite TCM cathode with ion-conductive 0.44LiBO2·0.56LiF as well as electron-conductive In2O5Sn (ITO) and reported that the resulting all-solid-state polymer battery using the copolymer P(EO/EH) of epichlorohydrin and ethylene oxide as the matrix produced a specific discharge capacity of 136 mAh g−1 with a surface capacity of up to 6 mAh cm−2 at 0.1 C after 20 cycles. Here, the researchers attributed these enhanced performances to the enhanced ionic and electronic percolation of the composite electrodes as well as the improved interfacial contact.
2.1.3 Polymers Containing Ethoxylated Sidechains
Polymers containing ethoxylated sidechains typically possess sufficient flexibility at the EO-based segment and favor ionic conductivity. For example, Ghamouss et al. [112] synthesized a solvent-free gel polymer matrix (P(MA475-DMA550)) through the free radical copolymerization of methacrylate-based oligomers in the presence of LiTFSI and ionic liquid RTIL and reported that the resultant free-standing GPE provided an ionic conductivity close to 4 × 10−4 S cm−1 at room temperature and that the corresponding LiNi1/3Mn1/3Co1/3O2/GPE/Li battery provided a good specific discharge capacity (118 mAh g−1 at 0.1 C and 79 mAh g−1 at 0.2 C). In another study, Shono et al. [111] conducted in-depth investigations into the capacity decay of polyether-based electrolytes by performing capacity fading analyses using pseudo-reference electrodes to measure the operation of a P(EO/MEEGE/AGE)-based solvent-free LIB with LiNi1/3Mn1/3Co1/3O2 as the cathode active material. Here, the researchers reported that the resultant cell achieved a capacity retention of 50% after 1000 charge/discharge cycles, and that the successive and irreversible loss of active lithium at the graphite anode is the main reason for capacity reduction, rather than the oxidation decomposition of the polyether-based electrolyte.
a–c Schematic illustration of the different interfacial chemistries of a bare Li metal and b Li metal with a grafted skin in a carbonate electrolyte. c Chemical structure of the polymer skin, composed of a cyclic ether group (pink) and a polycyclic main chain (blue) [113].
Copyright 2017, ACS
Polymers containing ethoxylated motifs have proven to be effective polymer matrixes for high energy density TCM-based LBs. However, insufficient ionic conductivity at room temperatures, low t Li + and inferior anti-oxidative abilities hinder application in high-voltage TCM-based LBs. Despite this, these polymer electrolytes can be further optimized through the introduction of functional groups with high oxidative resistances and strong coordination capabilities with transition metal ions (e.g., cyano or sulfone groups) as well as blending with advanced inorganic electrolytes with high ionic conductivities.
2.2 Polyacrylates
Reported polyacrylate-based electrolytes and corresponding electrochemical properties
Polymer matrix | Organic solvents/lithium salts/other additives | Ionic conductivity (S cm−1) | ESW versus Li (V) | Cell configuration | Voltage region versus Li (V) | Year | Ref. |
---|---|---|---|---|---|---|---|
PMMA/PVDF | EC–EMC–DMC/LiPF6 | 3.2 × 10−4 at 25 °C | 4.5 | LiNi1/3Mn1/3Co1/3O2/graphite | 3.0–4.3 | 2007 | [114] |
Poly(ethyl α-cyanoacrylate) | EC–DMC/LiPF6 | 2.54 × 10−3 at 25 °C | 4.7 | LiNi0.5Co0.2Mn0.3O2/graphite | 3.0–4.3 | 2015 | [115] |
PTAEP | EC–DMC/LiPF6 | – | – | LiNi1/3Mn1/3Co1/3O2/lithium metal | 2.8–4.6 | 2013 | [116] |
ipn-PEA | LiPF6 | 2.2 × 10−4 at 25 °C | 4.5 | LiNi0.5Co0.2Mn0.3O2/Li | 2.5–4.2 | 2016 | [117] |
P(PETEA) | EC–DEC–EMC/LiPF6 | 8.46 × 10−3 at 25 °C | – | LiNi0.8Co0.15Al0.05O2/graphite or graphite–Si/C | 2.75–4.2 | 2017 | [118] |
2.2.1 Polymethyl Methacrylate
Polymethyl methacrylate (PMMA) exhibits an excellent affinity to liquid electrolytes and can accommodate large uptake of liquid electrolytes. In addition, GPEs based on PMMA possess relatively good compatibility with lithium anodes. However, this polymer inherently suffers from insufficient mechanical strength, and to address this, Holze et al. [114] blended in poly(vinylidene fluoride) (PVDF) to enhance the mechanical properties of PMMA-based polymer matrixes and reported that after the absorption of liquid electrolytes, the blended polymer-based electrolyte provided an electrochemical stability window of 4.5 V and an ionic conductivity of 3.2 × 10−4 S cm−1 at 25 °C. Despite this, however, this GPE still suffered from inferior cycling stabilities, especially in the presence of lithium anodes.
2.2.2 Polycyanoacrylate
Polycyanoacrylate-based electrolytes can exhibit high anti-oxidative abilities and can be used as a polymer matrix to couple with high-voltage TCM-based cathodes. For example, our group [115] developed a composite electrolyte comprised of poly(ethyl α-cyanoacrylate) integrated with a poly(ethylene terephthalate) (PET) non-woven film by using a rigid-flexible coupling strategy to optimize the cycling ability of NCM-based LIBs at high operating voltages. Here, the resulting GPE provided a wide electrochemical stability window of 4.7 V and a good ionic conductivity (2.54 × 10−3 S cm−1 at room temperature), with the corresponding NCM/graphite full cell demonstrating excellent cycling stabilities with a capacity retention of 91% after 200 cycles (113 mAh g−1 at the 200th cycle) at 0.5 C. These results demonstrated that this type of GPEs can suppress the dissolution and diffusion of transition metal ions from the NCM cathode at high working potentials and as a result, polymer electrolytes using polycyanoacrylate as a matrix are potential candidates to match with TCM-based LBs at highly delithiated states. However, this polymer in TCM-based LBs still face challenges, such as poor stability in the presence of lithium anodes and insufficient mechanical properties.
2.2.3 Cross-Linked Polyacrylates
Cross-linked polymers can exhibit superior mechanical properties, and the mechanical strength of cross-linked polyacrylates can outperform linear polyacrylates. As a result, researchers have attempted to develop cross-linked polyacrylate-based polymer electrolytes for application in TCM-based LBs. For example, Lee et al. [116] synthesized a poly(tris(2-(acryloyloxy)ethyl) phosphate) (PTAEP)-based GPE using a direct UV-assisted coating method on an as-formed LiNi1/3Co1/3Mn1/3O2 cathode and reported that the synthesized GPE endowed the corresponding cell with superior capacity retention (84% vs. 73%) as compared with pristine cells at 1 C after 50 cycles with a charge cutoff voltage of 4.6 V. Here, the researchers reported that PTAEP can act as an ion-conductive protective film to suppress interfacial side reactions close to the cathode, and that the body of the phosphate groups in the polymer host can provide excellent safety properties as a flame retardant.
In another example, Guo et al. [117] developed a poly(ethoxylated trimethylolpropane triacrylate)/PEO (ipn-PEA)-based APE with a rigid-flexible network for room temperature LBs and reported an excellent ionic conductivity of 2.2 × 10−4 S cm−1 with a high mechanical strength of approximately 12 GPa, and reported that the corresponding cell using LiNi0.5Co0.2Mn0.3O2 as the cathode active material at a charge cutoff voltage of 4.3 V produced a specific capacity of 117 mAh g−1 after 200 cycles at 0.5 C. Here, the researchers reported that the interpenetrating polymer network with its robust mechanical strength not only suppressed the formation of lithium dendrites, but also improved interfacial stability during long-term cycling, providing significant insights into the relationship between interfacial compatibility and structural features of polymer matrixes in TCM-based LBs.
Electrochemical performances of the four types of full batteries. a–c Cycling performances at a 0.5 C/1 C at 25 °C, b 0.5 C/1 C at 45 °C and c 0.5 C/5 C at 25 °C. d Rate performances at 25 °C. e Gas generation of the batteries after 280 cycles at a temperature of 45 °C, and f after 200 cycles at a large discharge rate of 5 C. Here, NCA/liquid electrolyte/graphite, NCA/liquid electrolyte/(graphite–Si/C), NCA/PETEA-based GPE/graphite and NCA/PETEA-based GPE/(graphite–Si/C) are denoted as NLG, NLGS, NPG and NPGS, respectively [118].
Copyright 2017, RSC
Schematic illustrations of LBs during cycling: a LiNi0.8Co0.15Al0.05O2 (NCA)/liquid electrolyte/graphite and b NCA/PETEA-based gel polymer electrolyte/graphite [118].
Copyright 2017, RSC
Overall, polyacrylate-based electrolytes are promising candidates to replace conventional liquid electrolytes and resolve the safety concerns of TCM-based LBs. And among these, cross-linked polyacrylates demonstrate attractive properties such as the significant suppression of lithium dendrite formation and the reduction of liquid electrolyte decomposition, all of which can advance the future development of TCM-based LBs. However, side reactions between ester groups in polyacrylate polymer structures and lithium are prevalent, but may be optimized through the addition of film-forming additives on the anode such as fluoroethylene carbonate (FEC) and vinylene carbonate (VC) or through the introduction of organic or inorganic coatings onto the lithium metal anode.
2.3 Polycarbonates
Reported polycarbonate-based electrolytes and their electrochemical properties
Polymer matrix | Organic solvents/lithium salts/other additives | Ionic conductivity (S cm−1) | ESW versus Li (V) | Cell configuration | Voltage range versus Li (V) | Year | Ref. |
---|---|---|---|---|---|---|---|
XPEEC-1 | LiTFSI | 1.6 × 10−5 at 25 °C | 4.9 | LiNi0.6Mn0.2Co0.2O2/Li | 3.0–4.3 | 2017 | [119] |
P(OEGMA-CCMA) | EC–DMC/LiPF6 | 2.3 × 10−3 at 25 °C | 5.5 | LiNi1/3Mn1/3Co1/3O2/graphite | 2.8–4.2 | 2014 | [120] |
PVCA | EC–DMC/LiDFOB | 5.59 × 10−4 at 25 °C | 4.8 | LiNi0.6Mn0.2Co0.2O2/graphite | 3.0–4.2 | 2017 | [121] |
2.3.1 Poly(Ethylene Ether Carbonate)
PEEC-based electrolytes prepared through a ring-opening polymerization of ethylene carbonate can exhibit superior ionic conductivities and higher t Li + than those of PEO-based electrolytes. For example, Kim et al. [119] synthesized an amorphous polymer matrix (XPEEC-1) through a photocross-linking reaction between PEEC and tetraethyleneglycol diacrylate (TEGDA) and reported that the cross-linked polymer electrolyte possessed a high ionic conductivity of 1.6 × 10−5 S cm−1 at room temperature and a high t Li + of 0.40. In addition, the corresponding solid-state Li/LiNi0.6Co0.2Mn0.2O2 cell in this study delivered an initial discharge capacity of 141.4 mAh g−1 and good capacity retention (90.2% of the initial discharge capacity after 100 charge/discharge cycles at 0.1 C) at 25 °C, demonstrating potential application in all-solid-state TCM-based LBs under ambient temperatures.
2.3.2 Polycyclic Carbonate
Compared with linear carbonates, cyclic carbonates possess higher dielectric constants and therefore can dissociate lithium ions more efficiently. In addition, cyclic carbonate polymers can absorb large amounts of the liquid electrolyte and demonstrate higher mechanical strengths. In one study, Lex-Balducci et al. [120] synthesized a polymer matrix (P(OEGMA-CCMA)) by copolymerizing cyclic carbonate methacrylate (CCMA) with oligo (ethylene glycol) methyl ether methacrylate (OEGMA) and reported that the significant uptake of the liquid electrolyte in the resulting polymer host structure (550% of its own weight) resulted in a high ionic conductivity of 2.3 × 10−3 S cm−1 at 25 °C and that the corresponding LiNi1/3Co1/3Mn1/3O2/graphite full cell exhibited stable cycling characteristics with high discharge capacities (> 91% capacity retention after 80 cycles with a capacity of 122 mAh g−1) and Coulombic efficiencies (> 99.5%) at 0.5 C. In addition, the researchers also reported that the cyclic carbonate motif endowed the obtained polymer electrolyte with good mechanical stability. In another study, our group [121] developed a novel poly(vinylene carbonate) (PVCA)-based GPE through an in situ polymerization method that possessed a superior mechanical property of 2 GPa, outperforming most polymer electrolytes, and produced a high ionic conductivity of 5.59 × 10−4 S cm−1 and a wide electrochemical stability window of above 4.8 V versus Li at 25 °C. In addition, the corresponding LiNi0.6Co0.2Mn0.2O2/PVCA-based GPE/graphite cell demonstrated excellent safety and interfacial compatibility, indicating that PVCA-based electrolytes are potential alternatives for TCM-based LBs.
Polycarbonate-based electrolytes exhibit relatively strong coordination abilities to lithium ions and usually possess high ionic conductivities and polycyclic carbonates that endow electrolytes with high mechanical properties and the uptake of liquid electrolytes should be further explored as polymer matrixes for TCM-based LBs. However, polycarbonate-based electrolytes suffer from inferior chemical stabilities in the presence of lithium or basic cathodes, but this defect can be resolved through the introduction of atomic layer coatings onto the lithium anode.
2.4 Vinylidene Fluoride-Based Polymers
Reported VDF-based polymer-based electrolytes and their electrochemical properties
Polymer matrix | Organic solvents/lithium salts/other additives | Ionic conductivity (S cm−1) | ESW versus Li (V) | Cell configuration | Voltage range vs. Li (V) | Year | Ref. |
---|---|---|---|---|---|---|---|
PVDF | EC–PC/LiPF6 | – | – | LiNi1/3Mn1/3Co1/3O2/Li4Ti5O12 | 3.0–4.6 | 2007 | [123] |
P(VDF-HFP) | EC–DMC/LiPF6/TiO2 | 0.98 × 10−3 at 20 °C | – | LiNi1/3Mn1/3Co1/3O2/graphite | 2.8–4.2 | 2008 | [124] |
P(VDF-HFP) | BMP-TFSI/LiTFSI/VC | – | 4.8 | LiNi1/3Mn1/3Co1/3O2/Li | 3.0–4.3 | 2011 | [125] |
P(VDF-HFP) | MPPyrr-TFSA-PC/Li-TFSA | 1.23 × 10−3 at 20 °C | 5.0 | LiNi1/3Mn1/3Co1/3O2/graphite | 3.0–4.2 | 2013 | [126] |
P(VDF-HFP) | EC–DEC/LiPF6/SiO2 | 1.7 × 10−3 at 25 °C | – | LiNi1/3Mn1/3Co1/3O2/carbon | 3.0–4.5 | 2013 | [127] |
P(VDF-HFP) | EC–DMC/LiCF3SO3/Al2O3 | 7.1 × 10−3 at 25 °C | 4.9 | LiNi1/3Mn1/3Co1/3O2/Li | 2.5–4.6 | 2013 | [128] |
P(VDF-HFP) | EC–DEC/LiPF6/silica fillers | 3.78 × 10−3 at 30 °C | – | LiNi0.5Co0.2Mn0.3O2/Li | 2.5–4.3 | 2014 | [129] |
P(VDF-HFP) | EC–DMC/LiPF6/Al2O3 | 4.1 × 10−3 at 25 °C | 5.0 | LiNi1/3Mn1/3Co1/3O2/Li4Ti5O12 | 3.5–4.2 | 2017 | [130] |
P(VDF-TrFE) | EC–DEC/LiPF6 | – | – | Li1.2Mn0.54Ni0.13Co0.13O2/Li | 2.0–4.6 | 2017 | [131] |
P(VDF-HFP) | EC–EMC/LiPF6/TiO2 | – | – | LiMn0.8Ni0.1Co0.1O2/Li | 2.8–4.2 | 2018 | [132] |
2.4.1 Poly(Vinylidene Fluoride-Hexafluoropropylene)
P(VDF-HFP) contains an amorphous phase (-HFP) which can immobilize large amounts of the liquid electrolyte as well as a crystalline phase (-VDF) which can enhance mechanical properties, and compared with PVDF, P(VDF-HFP) exhibits decreased crystallinity and glass transition temperatures. Researchers have also reported that ionic liquids as additives can further optimize P(VDF-HFP)-based electrolyte performances due to desirable properties such as low flammability, excellent thermal stability, high ionic conductivity and more [125, 126]. In one study, Kim et al. [125] reported that a cell comprised of the LiNi1/3Mn1/3Co1/3O2/GPE containing P(VDF-HFP) and ionic liquid/Li delivered better cyclability than cells based on GPEs containing commercial liquid electrolytes (1.0 M LiPF6 in EC/DEC). In this study, it should be noted that additives such as VC, FEC and EC were also applied to stabilize the SEI layer and alleviate the reductive breakdown of ionic liquids, leading to good compatibility between the GPE and the lithium anode. In another study, Hofmann et al. [126] reported that ionic liquids play an important role in promoting the cycling capability of TCM-based cells.
Aside from ionic liquids, the addition of inorganic fillers can also enhance the cell performance of P(VDF-HFP)-based GPEs. For example, Wu et al. [124] synthesized a macroporous nanocomposite polymer membrane comprised of P(VDF-HFP) and TiO2 through an in situ hydrolysis of Ti(OC4H9) and reported that the resulting GPE demonstrated enhanced ionic conductivities (9.8 × 10−4 vs. 1.4 × 10−4 S cm−1 at 20 °C) and entrapment of liquid electrolytes (125 vs. 56 wt%), as well as decreased crystallinity (a melting point of 146.6 vs. 142.2 °C) and better rate capabilities (> 10 mAh g−1 higher at 0.5–5 C) than the GPE without TiO2. In a further study, Kim et al. [127] prepared another composite GPE with P(VDF-HFP) and core–shell SiO2(Li+) nanoparticles which also presented a high ionic conductivity of 1.7 × 10−3 S cm−1. Here, the composite membrane-assembled LiNi1/3Co1/3Mn1/3O2/carbon cell demonstrated a superior capacity retention of 95% after 100 cycles at 0.5 C and a better rate capability of 167 mAh g−1 at 5 C rate as compared with that of a pristine P(VDF-HFP)-based membrane. Furthermore, other inorganic fillers including mesoporous modified-silica fillers and Al2O3 have also been blended with P(VDF-HFP), and the obtained composite films have also been reported to exhibit superior mechanical and electrochemical properties as compared with TCM cathodes [128, 129, 130]. In addition, Wang et al. [132] recently studied a Li-eliminating composite GPE to avoid electrical short circuiting in which the Li-eliminating GPE was prepared by coating a porous P(VDF-HFP) polymer matrix embedded with TiO2 nanoparticles onto the surface of a conventional Celgard separator which is facing the cathode. Here, the researchers reported that the TiO2-containing composite system can react with lithium dendrites penetrating through the separator during cycling and significantly decrease the short circuiting of the resulting Cu/Li-eliminating composite GPE/Li cell as compared with a pristine cell using a pristine polypropylene (PP) separator, allowing the corresponding LiNi0.8Mn0.1Co0.1O2/Li LB to demonstrate better capacity retentions (> 90% vs. 63% after 155 cycles at 1 C) and rate capabilities (approximately 115 vs. 49 mAh g−1 at 11 C).
2.4.2 Poly(Difluoroethylene-Trifluoroethylene)
Kong et al. [131] reported that poly(difluoroethylene-trifluoroethylene (P(VDF-TrFE)) possesses considerable electronic and ionic conductivities and can be coated onto Li1.2Mn0.54Ni0.13Co0.13O2 cathode surfaces to stabilize the cathode active material and mitigate unwanted interfacial side reactions close to the cathode. In their study, the researchers reported that their β-P(VDF-TrFE)-coated sample provided a high initial capacity of 262.8 mAh g−1 at 0.1 C, a superior capacity retention of 90.8% after 100 cycles at 0.1 C and a high rate capability of 116.8 mAh g−1 at 5 C, which exceeded the performance of samples with no coating, demonstrating that conductive P(VDF-TrFE) coatings have potential for practical application in TCM-based LBs.
The addition of ionic liquids or inorganic fillers into PVDF-HFP matrixes can improve the performance of resulting cells, and compared with ionic liquids, inorganic filler-based composite gel electrolytes demonstrate better mechanical and electrochemical properties and higher rate performances. Here, VDF-based copolymers as gel polymer matrixes are promising in TCM-based cell applications due to inherent advantages such as outstanding nonflammability. Despite these properties, however, these polymers suffer from high production costs and insufficient stability with lithium.
2.5 Polyacrylonitrile
Reported PAN-based electrolytes and their electrochemical properties
Polymer matrix | Organic solvents/lithium salts/other additives | Ionic conductivity (S cm−1) | ESW versus Li (V) | Cell configuration | Voltage range versus Li (V) | Year | Ref. |
---|---|---|---|---|---|---|---|
PAN | PC/LiPF6 | 8 × 10−3 at 25 °C | 5.5 | LiNi0.8Co0.16Al0.04O2/lithium metal | 3.5–5.1 | 2002 | [133] |
P(MA-SiO2-TEGDA)/PAN | EC–EMC/LiPF6 | 1.8 × 10−3 at 25 °C | – | LiNi1/3Co1/3Mn1/3O2/graphite | 3.0–4.5 | 2016 | [134] |
P(MA-SiO2-TEGDA)/PAN | EC–EMC–DEC/LiPF6/FEC | 2.1 × 10−3 at 25 °C | – | LiNi0.6Co0.6Mn0.2O2/graphite | 2.6–4.3 | 2017 | [135] |
Overall, PAN-based polymer electrolytes possess several merits such as high oxidative resistances. However, these electrolytes suffer from inferior compatibility with lithium metal anodes and poor mechanical strength. Here, interfacial optimization close to lithium metal anodes can be achieved through the copolymerization or introduction of inorganic or organic additives as well as polymer blending.
2.6 Aromatic Polymers
Reported aromatic polymer-based electrolytes and their electrochemical properties
Polymer matrix | Organic solvents/lithium salts/other additives | Ionic conductivity (S cm−1) | ESW versus Li (V) | Cell configuration | Voltage range versus Li (V) | Year | Ref. |
---|---|---|---|---|---|---|---|
P(PMDA-ODA) | EC–DMC/LiPF6 | 1.5 × 10−4 | – | LiNi1/3Co1/3Mn1/3O2/graphite | 2.9–4.8 | 2012 | [136] |
P(BPADA-pPD) | EC–DMC/LiPF6/VC | – | 5.0 | LiNi0.6Co0.6Mn0.2O2/graphite | 2.7–4.2 | 2017 | [137] |
PPy | EC–EMC/LiPF6 | – | – | LiNi1/3Co1/3Mn1/3O2/Li metal | 2.8–4.6 | 2011 | [139] |
PPy | EC–DMC–EMC/LiPF6 | – | – | LiNi0.8Co0.1Mn0.1O2/Li metal | 2.8–4.3 | 2014 | [140] |
PPy | EC–DMC/LiPF6 | – | – | Li1.2Ni0.54Co0.13Mn0.13O2/Li metal | 2.0–4.8 | 2013 | [141] |
Zirconium-doped PPy | EC–DMC–EMC/LiPF6 | – | – | Li1.2Ni0.5Co0.2Mn0.3O2/Li metal | 3.0-4.6 | 2016 | [142] |
2.6.1 Polyimide
PI polymers can generally be prepared through the thermal imidization of diamine and dianhydride monomers and possess rigid polymer backbones and high electrochemical stability windows. In addition, these polymers can effectively prevent close contact between TCM cathodes and liquid electrolytes, and therefore suppress CEI irreversible side reactions. Lee et al. [136] were the first to propose a surface modification strategy based on P(PMDA-ODA)-directed nanoscale (about 10 nm) skins which can significantly improve the performance and thermal stability of high-voltage TCM-based LIBs. In their study, the researchers wrapped a LiNi1/3Co1/3Mn1/3O2 surface with a nanoscale layer of polyimide and reported good cycling performances (66% capacity retention after the 50th cycle) at high operating voltages (2.8–4.8 V vs. Li). However, the researchers also reported that the ionic conductivity of the PI film with the uptake of the liquid electrolyte was relatively low (1.5 × 10−4 S cm−1) and resulted in poor LIB rate performances with the PI coating. In another study, l’Abee et al. [137] reported that their PI-based separator (P(BPADA-pPD)) exhibited superior total porosity (up to 71%), excellent anti-oxidative ability (5 V vs. Li), high electrolyte uptake (> 300 wt%), and high dimensional stability (up to 220 °C), with the corresponding NMC/graphite pouch cell using the highly porous separator demonstrating enhanced capacity retention (89.3% vs. 81.9%) and slightly lower Coulombic efficiency (99.77% vs. 99.91%) as compared with their polyolefin reference after 1000 (1 C/2 C) cycles. Overall, although these results suggest that polyimides are promising for TCM-based high-voltage LBs, P(BPADA-pPD) solid films suffer from long soaking times (up to 21 days) and unsatisfactory rate performances, which can be optimized through the introduction of F or CF3 in the polymer host or the addition of highly polar additives such as ionic liquids in terms of the liquid electrolyte. In addition, the synthesis of PI is complex and expensive and is unsuitable for large-scale industrial production applications.
2.6.2 Polypyrrole
Different from PI-based protective layers, PPy coatings possess superior electronic conductivities (on the order of 102 S cm−1 [138]) due to an inherently conjugated and relatively coplanar backbone. In addition, these coatings are cost-effective, easily prepared and are electrochemically active for lithium ion deposition and stripping (theoretical capacity of 72 mAh g−1) at 2.0–4.5 V versus Li, allowing PPy coatings to act as both a conductive film and an organic cathode material. A PPy-coated TCM cathode was first prepared by Ren et al. [139] in 2011 in which the researchers reported that the obtained cell using lithium metal as the anode produced a specific discharge capacity of 182 mAh g−1 after 50 cycles at 0.1 C and voltage limits of 2.8–4.6 V, outperforming that of bare cathode-based LBs (only 134 mAh g−1). Wang et al. [140] also reported that PPy-coated cathodes possessed notably enhanced cycling stabilities (69.8% vs 44.5% capacity retention after 100 cycles at 2 C rate and 60 °C) and rate capabilities (132.4 vs 115.2 mAh g−1 at 10 C under ambient temperature) as compared with pristine materials. Here, the researchers suggested that these enhanced battery performances can be attributed to the beneficial characteristics of the PPy coating, including the increased electron conductivity and the suppression of unwanted side reactions between the highly delithiated Li1Ni0.8Co0.1Mn0.1O2 and the electrolyte. In another studies, PPy has also been used to coat high-voltage Li1.2Ni0.54Co0.13Mn0.13O2 and Li1Ni0.5Co0.2Mn0.3O2 and has been reported to enhance cell cycling performances and rate capabilities [141, 142]. These results indicate that PPy-based surface modifications on TCM cathodes can enhance cell performances such as rate performances, and are promising for practical application.
Overall, aromatic polymers can prevent contact between TCM cathodes and liquid electrolytes, and significantly optimize cell performances under high voltages or temperatures. As for PI coating-based cell systems, they usually require complex and expensive processing steps and possess unsatisfactory rate capabilities. Alternatively, PPy systems possess advantageous characteristics such as easy preparation, low costs and high electronic conductivity, allowing corresponding cells to possess superior rate performances. In general, cells with PPy-coated cathodes are not considered to be GPE cells due to the low uptake of liquid electrolytes by PPy. However, this disadvantage can be overcome through the blending with other polymers or through structural modifications to enhance liquid electrolyte immobilization.
3 Conclusion and Perspectives
- (1)
Polymers with EO motifs: Polymers with EO motifs have been mainly employed in APE LBs based on TCM due to their ability to coordinate and dissociate lithium salts as well as their flexible backbone and good mechanical strength. However, these EO motif polymers suffer from low ionic conductivities at room temperature, low t Li + and inferior anti-oxidative abilities, and further research should focus on bulk modifications through the introduction of functional groups with high oxidative resistances and strong coordination capabilities with transition metal ions, as well as blending with advanced inorganic electrolytes.
- (2)
Polyacrylates: Polyacrylates-based GPEs are inexpensive, possess good interfacial compatibility, and are promising alternatives to replace conventional liquid electrolytes and resolve associated safety hazards in TCM-based LBs. This is especially true for cross-linked polyacrylates which possess attractive advantages such as the suppression of lithium dendrite formation and liquid electrolyte decomposition, and are promising in TCM-based LBs. As for the successful application of cross-linked polyacrylate-based electrolytes, future development should focus on the modification of lithium anode surfaces to enhance interfacial compatibility.
- (3)
Polycarbonates: Polycarbonates-based GPEs are another type of polymer electrolytes in which polycyclic carbonate-based GPEs possess good mechanical properties and excellent immobilization of liquid electrolytes and deserve to be further developed for TCM-based LBs.
- (4)
VDF-based copolymers: VDF-based copolymers possess desirable internal strengths such as outstanding nonflammability characteristics, making them good choices for high safety LBs based on TCM. The drawbacks of this type of polymer electrolyte, including insufficient stability with lithium anodes, can be resolved through the modification of lithium metal anode surfaces, the introduction of organic/inorganic additives, and the use of solvent-in-salt systems.
- (5)
Polyacrylonitrile (PAN)-based GPEs: PAN-based GPEs possess superior anti-oxidative abilities, high ionic conductivities and thermal stability, and have great potential as hosts to match with high-voltage TCM-based cathodes. However, these polymers possess inferior interfacial compatibility with lithium metal anodes, which can be improved through copolymerization, introduction of inorganic or organic additives, and polymer blending.
- (6)
Aromatic polymers: Compared with the above-mentioned conventional polymer hosts, aromatic polymers can build robust barriers between TCM cathodes and liquid electrolytes and suppress undesirable side reactions close to the cathode. These characteristics endow resultant cells with significantly improved electrochemical performances under high-voltage or high-temperature conditions. And of the various aromatic polymers, cells with PPy-coated cathodes cannot be considered as a GPE due to the low uptake of the liquid electrolyte and should be used together with other polymer matrixes to obtain excellent electrochemical properties in TCM-based LBs.
- (1)
Because organic solvents are unavoidable in GPEs, leading to potential safety risks, it is necessary to develop efficient flame-retardant GPEs to ensure application safety in LBs employing conventional GPEs. In addition, it is of great importance to evaluate the thermal safety of polymer electrolyte-based LBs based on TCM through Accelerating Rate Calorimeter.
- (2)
More efforts need to be devoted to the investigation of interfacial properties between electrodes and polymer electrolytes by using in situ technologies, especially the evolution of CEI layers.
- (3)
The development of intelligent polymer electrolytes with intelligent responsive properties (i.e., anti-thermal or shock-proof GPEs) is a promising developmental route.
- (4)
Novel preparation methods (apart from solution-casting methods and in situ polymerization methods) of polymer electrolytes need to be developed.
- (5)
Focuses should be placed on the development of flexible and stretchable polymer electrolytes as wearable and smart device technologies become more mature.
- (6)
Because polymer matrixes containing interfacial film-forming motifs possess excellent compatibility with corresponding electrodes [143], this is an effective method to optimize interfacial properties between polymer electrolytes and electrodes.
Notes
Acknowledgements
This study was financially supported by the National Natural Science Fund for Distinguished Young Scholars (51625204), the Key Research and Development Plan of Shandong Province P. R. China (2017GGX40119), the National Natural Science Foundation of China (51703236 and 51803230), the Youth Innovation Promotion Association of CAS (2016193), and the National Key R&D Program of China (Grant No. 2018YFB0104300).
References
- 1.Pan, S., Ren, J., Fang, X., et al.: Integration: an effective strategy to develop multifunctional energy storage devices. Adv. Energy Mater. 6, 1501867 (2016)CrossRefGoogle Scholar
- 2.Kim, J., Kumar, R., Bandodkar, A.J., et al.: Advanced materials for printed wearable electrochemical devices: a review. Adv. Electron. Mater. 3, 1600260 (2017)CrossRefGoogle Scholar
- 3.Cheng, X.L., Pan, J., Zhao, Y., et al.: Gel polymer electrolytes for electrochemical energy storage. Adv. Energy Mater. 8, 1702184 (2018)CrossRefGoogle Scholar
- 4.Vandepaer, L., Cloutier, J., Amor, B.: Environmental impacts of lithium metal polymer and lithium-ion stationary batteries. Renew. Sustain. Energy Rev. 78, 46–60 (2017)CrossRefGoogle Scholar
- 5.Sun, C.W., Liu, J., Gong, Y.D., et al.: Recent advances in all-solid-state rechargeable lithium batteries. Nano Energy 33, 363–386 (2017)CrossRefGoogle Scholar
- 6.Manthiram, A., Yu, X.W., Wang, S.F.: Lithium battery chemistries enabled by solid-state electrolytes. Nat. Rev. Mater. 2, 16103 (2017)CrossRefGoogle Scholar
- 7.Jiang, C., Li, H.Q., Wang, C.L.: Recent progress in solid-state electrolytes for alkali-ion batteries. Sci. Bull. 62, 1473–1490 (2017)CrossRefGoogle Scholar
- 8.Zhou, G.M., Li, F., Cheng, H.M.: Progress in flexible lithium batteries and future prospects. Energy Environ. Sci. 7, 1307–1338 (2014)CrossRefGoogle Scholar
- 9.Liu, J., Xu, J.Y., Lin, Y., et al.: All-solid-state lithium ion battery: research and industrial prospects. Acta Chim. Sin. 71, 869–878 (2013)CrossRefGoogle Scholar
- 10.Peng, H.J., Huang, J.Q., Zhang, Q.: A review of flexible lithium–sulfur and analogous alkali metal–chalcogen rechargeable batteries. Chem. Soc. Rev. 46, 5237–5288 (2017)CrossRefPubMedGoogle Scholar
- 11.Saha, P., Datta, M.K., Velikokhatnyi, O.I., et al.: Rechargeable magnesium battery: current status and key challenges for the future. Prog. Mater. Sci. 66, 1–86 (2014)CrossRefGoogle Scholar
- 12.Hueso, K.B., Palomares, V., Armand, M., et al.: Challenges and perspectives on high and intermediate-temperature sodium batteries. Nano Res. 10, 4082–4114 (2017)CrossRefGoogle Scholar
- 13.Che, H.Y., Chen, S.L., Xie, Y.Y., et al.: Electrolyte design strategies and research progress for room-temperature sodium-ion batteries. Energy Environ. Sci. 10, 1075–1101 (2017)CrossRefGoogle Scholar
- 14.Bloomberg New Energy Finance, Electric Vehicle Outlook 2017 (2017). https://about.bnef.com. Accessed 1 Jan 2019
- 15.Choi, J.W., Aurbach, D.: Promise and reality of post-lithium-ion batteries with high energy densities. Nat. Rev. Mater. 1, 16013 (2016)CrossRefGoogle Scholar
- 16.Li, J., Zhu, J., Li, Q., Huang, J., Tao, X.: Development of electrode materials for lithium ion battery with high energy density. New chem. Mater. 43, 15–16 (2015)Google Scholar
- 17.Yang, S., Ren, W., Chen, J.: Li-rich oxides cathode materials: towards a new generation of lithium-ion batteries with high energy density. Mater. Rev. 31, 1–10 (2017)Google Scholar
- 18.Liu, J.Y., Li, X.X., Huang, J.R., et al.: Three-dimensional graphene-based nanocomposites for high energy density Li-ion batteries. J. Mater. Chem. A 5, 5977–5994 (2017)CrossRefGoogle Scholar
- 19.Li, J., Du, Z., Ruther, R.E., et al.: Toward low-cost, high-energy density, and high-power density lithium-ion batteries. JOM 69, 1484–1496 (2017)CrossRefGoogle Scholar
- 20.Ma, J., Hu, P., Cui, G., et al.: Surface and interface issues in spinel LiNi0.5Mn1.5O: insights into a potential cathode material for high energy density lithium ion batteries. Chem. Mater. 28, 3578–3606 (2016)CrossRefGoogle Scholar
- 21.Placke, T., Kloepsch, R., Duehnen, S., et al.: Lithium ion, lithium metal, and alternative rechargeable battery technologies: the odyssey for high energy density. J. Solid State Electrochem. 21, 1939–1964 (2017)CrossRefGoogle Scholar
- 22.Radin, M.D., Hy, S., Sina, M., et al.: Narrowing the gap between theoretical and practical capacities in li-ion layered oxide cathode materials. Adv. Energy Mater. 7, 1602888 (2017)CrossRefGoogle Scholar
- 23.Myung, S.T., Maglia, F., Park, K.J., et al.: Nickel-rich layered cathode materials for automotive lithium-ion batteries: achievements and perspectives. ACS Energy Lett. 2, 196–223 (2017)CrossRefGoogle Scholar
- 24.Ding, Y., Mu, D., Wu, B., et al.: Recent progresses on nickel-rich layered oxide positive electrode materials used in lithium-ion batteries for electric vehicles. Appl. Energy 195, 586–599 (2017)CrossRefGoogle Scholar
- 25.Manthiram, A., Knight, J.C., Myung, S.T., et al.: Nickel-rich and lithium-rich layered oxide cathodes: progress and perspectives. Adv. Energy Mater. 6, 1501010 (2016)CrossRefGoogle Scholar
- 26.Liu, W., Oh, P., Liu, X., et al.: Nickel-rich layered lithium transition-metal oxide for high-energy lithium-ion batteries. Angew. Chem. Int. Ed. Engl. 54, 4440–4457 (2015)CrossRefPubMedGoogle Scholar
- 27.Zhao, X., Wang, J., Dong, X., et al.: Structure design and performance of LiNixCoyMn1−x−yO2 cathode materials for lithium-ion batteries: a review. J. Chin. Chem. Soc.-TAIP 61, 1071–1083 (2014)CrossRefGoogle Scholar
- 28.Chen, J.: Recent progress in advanced materials for lithium ion batteries. Materials 6, 156–183 (2013)CrossRefPubMedPubMedCentralGoogle Scholar
- 29.Xu, B., Qian, D., Wang, Z., et al.: Recent progress in cathode materials research for advanced lithium ion batteries. Mater. Sci. Eng. R 73, 51–65 (2012)CrossRefGoogle Scholar
- 30.Kraytsberg, A., Ein-Eli, Y.: Higher, stronger, better… a review of 5 V cathode materials for advanced lithium-ion batteries. Adv. Energy Mater. 2, 922–939 (2012)CrossRefGoogle Scholar
- 31.Yabuuchi, N., Makimura, Y., Ohzuku, T.: Solid-state chemistry and electrochemistry of LiCo1/3Ni1/3Mn1/3O2 for advanced lithium-ion batteries III. Rechargeable capacity and cycleability. J. Electrochem. Soc. 154, A314–A321 (2007)CrossRefGoogle Scholar
- 32.Ohzuku, T., Brodd, R.J.: An overview of positive-electrode materials for advanced lithium-ion batteries. J. Power Sources 174, 449–456 (2007)CrossRefGoogle Scholar
- 33.Nitta, N., Wu, F., Lee, J.T., et al.: Li-ion battery materials: present and future. Mater. Today 18, 252–264 (2015)CrossRefGoogle Scholar
- 34.Ben Yahia, H., Shikano, M., Kobayashi, H.: Phase transition mechanisms in LixCoO2 (0.25\(\leqslant\) x \(\leqslant\) 1) based on group–subgroup transformations. Chem. Mater. 25, 3687–3701 (2013)CrossRefGoogle Scholar
- 35.Chen, Z.H., Lu, Z.H., Dahn, J.R.: Staging phase transitions in LixCoO2. J. Electrochem. Soc. 149, A1604–A1609 (2002)CrossRefGoogle Scholar
- 36.Van der Ven, A., Aydinol, M.K., Ceder, G., et al.: First-principles investigation of phase stability in LixCoO2. Phys. Rev. B 58, 2975–2987 (1998)CrossRefGoogle Scholar
- 37.Wang, L., Maxisch, T.M., Ceder, G.: A first-principles approach to studying the thermal stability of oxide cathode materials. Chem. Mater. 19, 543–552 (2007)CrossRefGoogle Scholar
- 38.Noh, H.J., Youn, S., Yoon, C.S., et al.: Comparison of the structural and electrochemical properties of layered Li[NixCoyMnz]O2 (x = 1/3, 0.5, 0.6, 0.7, 0.8 and 0.85) cathode material for lithium-ion batteries. J. Power Sources 233, 121–130 (2013)CrossRefGoogle Scholar
- 39.Bak, S.M., Hu, E., Zhou, Y., et al.: Structural changes and thermal stability of charged LiNixMnyCozO2 cathode materials studied by combined in situ time-resolved XRD and mass spectroscopy. ACS Appl. Mater. Interfaces. 6, 22594–22601 (2014)CrossRefPubMedGoogle Scholar
- 40.Konishi, H., Yoshikawa, M., Hirano, T.: The effect of thermal stability for high-Ni-content layer-structured cathode materials, LiNi0.8Mn0.1−xCo0.1MoxO2 (x = 0, 0.02, 0.04). J. Power Sources 244, 23–28 (2013)CrossRefGoogle Scholar
- 41.Feng, X., Ouyang, M., Liu, X., et al.: Thermal runaway mechanism of lithium ion battery for electric vehicles: a review. Energy Storage Mater. 10, 246–267 (2018)CrossRefGoogle Scholar
- 42.Li, W., Song, B., Manthiram, A.: high-voltage positive electrode materials for lithium-ion batteries. Chem. Soc. Rev. 46, 3006–3059 (2017)CrossRefPubMedGoogle Scholar
- 43.Xia, L., Yu, L., Hu, D., et al.: Research progress and perspectives on high voltage, flame retardant electrolytes for lithium-ion batteries. Acta Chim. Sin. 75, 1183–1195 (2017)CrossRefGoogle Scholar
- 44.Tan, S., Ji, Y.J., Zhang, Z.R., et al.: Recent progress in research on high-voltage electrolytes for lithium-ion batteries. ChemPhysChem 15, 1956–1969 (2014)CrossRefPubMedGoogle Scholar
- 45.Xu, K.: Electrolytes and interphases in Li-ion batteries and beyond. Chem. Rev. 114, 11503–11618 (2014)CrossRefPubMedGoogle Scholar
- 46.Choi, N.S., Han, J.G., Ha, S.Y., et al.: Recent advances in the electrolytes for interfacial stability of high-voltage cathodes in lithium-ion batteries. RSC Adv. 5, 2732–2748 (2015)CrossRefGoogle Scholar
- 47.Zhang, L., Ma, Y., Du, C., et al.: Research on the high-voltage electrolyte for lithium ion batteries. Prog. Chem. 26, 553–559 (2014)Google Scholar
- 48.Liu, K., Pei, A., Lee, H.R., et al.: Lithium metal anodes with an adaptive “solid–liquid” interfacial protective layer. J. Am. Chem. Soc. 139, 4815–4820 (2017)CrossRefPubMedGoogle Scholar
- 49.Komaba, S., Kumagai, N., Kataoka, Y.: Influence of manganese (II), cobalt (II), and nickel (II) additives in electrolyte on performance of graphite anode for lithium-ion batteries. Electrochim. Acta 47, 1229–1239 (2002)CrossRefGoogle Scholar
- 50.Zheng, J., Yan, P., Zhang, J., et al.: Suppressed oxygen extraction and degradation of LiNixMnyCozO2 cathodes at high charge cut-off voltages. Nano Res. 10, 4221–4231 (2017)CrossRefGoogle Scholar
- 51.Wang, Z., Lu, H.Q., Yin, Y.P., et al.: FePO4-coated Li[Li0.2Ni0.13Co0.13Mn0.54]O2 with improved cycling performance as cathode material for Li-ion batteries. Rare Met. 36, 899–904 (2017)CrossRefGoogle Scholar
- 52.Min, K., Park, K., Park, S.Y., et al.: Improved electrochemical properties of LiNi0.91Co0.06Mn0.03O2 cathode material via Li-reactive coating with metal phosphates. Sci. Rep. 7, 7151 (2017)CrossRefPubMedPubMedCentralGoogle Scholar
- 53.McNulty, D., Geaney, H., O’Dwyer, C.: Carbon-coated honeycomb Ni–Mn–Co–O inverse opal: a high capacity ternary transition metal oxide anode for Li-ion batteries. Sci. Rep. 7, 42263 (2017)CrossRefPubMedPubMedCentralGoogle Scholar
- 54.He, L., Xu, J., Han, T., et al.: SmPO4-coated Li1.2Mn0.54Ni0.13Co0.13O2 as a cathode material with enhanced cycling stability for lithium ion batteries. Ceram. Int. 43, 5267–5273 (2017)CrossRefGoogle Scholar
- 55.Xie, Y., Gao, D., Zhang, L.L., et al.: CeF3-modified LiNi1/3CO1/3Mn1/3O2 cathode material for high-voltage Li-ion batteries. Ceram. Int. 42, 14587–14594 (2016)CrossRefGoogle Scholar
- 56.Loeffler, N., Kim, G.T., Mueller, F., et al.: In situ coating of Li[Ni0.33Mn0.33Co0.33]O2 particles to enable aqueous electrode processing. Chemsuschem 9, 1112–1117 (2016)CrossRefPubMedGoogle Scholar
- 57.Wang, L., Ma, Y., Li, Q., et al.: 1,3,6-Hexanetricarbonitrile as electrolyte additive for enhancing electrochemical performance of high voltage Li-rich layered oxide cathode. J. Power Sources 361, 227–236 (2017)CrossRefGoogle Scholar
- 58.Su, C.C., He, M., Redfern, P.C., et al.: Oxidatively stable fluorinated sulfone electrolytes for high voltage high energy lithium-ion batteries. Energy Environ. Sci. 10, 900–904 (2017)CrossRefGoogle Scholar
- 59.Hilbig, P., Ibing, L., Wagner, R., et al.: Ethyl methyl sulfone-based electrolytes for lithium ion battery applications. Energies 10, 1312 (2017)CrossRefGoogle Scholar
- 60.Tu, W., Xing, L., Xia, P., et al.: Dimethylacetamide as a film-forming additive for improving the cyclic stability of high voltage lithium-rich cathode at room and elevated temperature. Electrochim. Acta 204, 192–198 (2016)CrossRefGoogle Scholar
- 61.Liao, X., Zheng, X., Chen, J., et al.: Tris(trimethylsilyl)phosphate as electrolyte additive for self-discharge suppression of layered nickel cobalt manganese oxide. Electrochim. Acta 212, 352–359 (2016)CrossRefGoogle Scholar
- 62.Mai, S., Xu, M., Liao, X., et al.: Tris(trimethylsilyl) phosphite as electrolyte additive for high voltage layered lithium nickel cobalt manganese oxide cathode of lithium ion battery. Electrochim. Acta 147, 565–571 (2014)CrossRefGoogle Scholar
- 63.Todorov, Y.M., Fujii, K., Yoshimoto, N., et al.: Ion-solvation structure and battery electrode characteristics of nonflammable organic electrolytes based on tris(trifluoroethyl)phosphate dissolving lithium salts. Phys. Chem. Chem. Phys. 19, 31085–31093 (2017)CrossRefPubMedGoogle Scholar
- 64.Zeng, Z., Wu, B., Xiao, L., et al.: Safer lithium ion batteries based on nonflammable electrolyte. J. Power Sources 279, 6–12 (2015)CrossRefGoogle Scholar
- 65.Feng, J.K., Sun, X.J., Ai, X.P., et al.: Dimethyl methyl phosphate: a new nonflammable electrolyte solvent for lithium-ion batteries. J. Power Sources 184, 570–573 (2008)CrossRefGoogle Scholar
- 66.Xu, G., Pang, C., Chen, B., et al.: Prescribing functional additives for treating the poor performances of high-voltage (5 V-class) LiNi0.5Mn1.5O4/MCMB Li-ion batteries. Adv. Energy Mater. 8, 1701398 (2018)CrossRefGoogle Scholar
- 67.Xu, G., Liu, Z., Zhang, C., et al.: Strategies for improving the cyclability and thermo-stability of LiMn2O4-based batteries at elevated temperatures. J. Mater. Chem. A 3, 4092–4123 (2015)CrossRefGoogle Scholar
- 68.Pang, C., Xu, G., An, W., et al.: Three-component functional additive in a LiPF6-based carbonate electrolyte for a high-voltage LiCoO2/graphite battery system. Energy Technol. 5, 1979–1989 (2017)CrossRefGoogle Scholar
- 69.Zeng, Z., Murugesan, V., Han, K.S., et al.: Non-flammable electrolytes with high salt-to-solvent ratios for Li-ion and Li-metal batteries. Nat. Energy 3, 674–681 (2018)CrossRefGoogle Scholar
- 70.Shi, P., Zheng, H., Liang, X., et al.: A highly concentrated phosphate-based electrolyte for high-safety rechargeable lithium batteries. Chem. Commun. 54, 4453–4456 (2018)CrossRefGoogle Scholar
- 71.Suo, L., Xue, W., Gobet, M., et al.: Fluorine-donating electrolytes enable highly reversible 5-V-class Li metal batteries. PNAS 115, 1156–1161 (2018)CrossRefPubMedGoogle Scholar
- 72.Jiao, S., Ren, X., Cao, R., et al.: Stable cycling of high-voltage lithium metal batteries in ether electrolytes. Nat. Energy 3, 739–746 (2018)CrossRefGoogle Scholar
- 73.Alvarado, J., Schroeder, M.A., Zhang, M., et al.: A carbonate-free, sulfone-based electrolyte for high-voltage Li-ion batteries. Mater. Today 21, 341–353 (2018)CrossRefGoogle Scholar
- 74.Wang, J., Yamada, Y., Sodeyama, K., et al.: Fire-extinguishing organic electrolytes for safe batteries. Nat. Energy 3, 22–29 (2017)CrossRefGoogle Scholar
- 75.Shiga, T., Kato, Y., Kondo, H., et al.: Self-extinguishing electrolytes using fluorinated alkyl phosphates for lithium batteries. J. Mater. Chem. A 5, 5156–5162 (2017)CrossRefGoogle Scholar
- 76.Wang, J., Yamada, Y., Sodeyama, K., et al.: Superconcentrated electrolytes for a high-voltage lithium-ion battery. Nat. Commun. 7, 12032 (2016)CrossRefPubMedPubMedCentralGoogle Scholar
- 77.Abbrent, S., Greenbaum, S.: Recent progress in NMR spectroscopy of polymer electrolytes for lithium batteries. Curr. Opin. Colloid Interface Sci. 18, 228–244 (2013)CrossRefGoogle Scholar
- 78.Arya, A., Sharma, A.L.: Polymer electrolytes for lithium ion batteries: a critical study. Ionics 23, 497–540 (2017)CrossRefGoogle Scholar
- 79.Ahmad, S.: Polymer electrolytes: characteristics and peculiarities. Ionics 15, 309–321 (2009)CrossRefGoogle Scholar
- 80.Zhang, Q.Q., Liu, K., Ding, F., et al.: Recent advances in solid polymer electrolytes for lithium batteries. Nano Res. 10, 4139–4174 (2017)CrossRefGoogle Scholar
- 81.Varshney, P.K., Gupta, S.: Natural polymer-based electrolytes for electrochemical devices: a review. Ionics 17, 479–483 (2011)CrossRefGoogle Scholar
- 82.Meyer, W.H.: Polymer electrolytes for lithium-ion batteries. Adv. Mater. 10, 439–448 (1998)CrossRefPubMedGoogle Scholar
- 83.Takeda, Y., Imanishi, N., Yamamoto, O.: Developments of the advanced all-solid-state polymer electrolyte lithium secondary battery. Electrochemistry 77, 784–797 (2009)CrossRefGoogle Scholar
- 84.Li, J., Ma, C., Chi, M., et al.: Solid electrolyte: the key for high-voltage lithium batteries. Adv. Energy Mater. 5, 1401408 (2015)CrossRefGoogle Scholar
- 85.Hu, Z., Zhang, S., Dong, S., et al.: Poly(ethyl α-cyanoacrylate)-based artificial solid electrolyte interphase layer for enhanced interface stability of Li metal anodes. Chem. Mater. 29, 4682–4689 (2017)CrossRefGoogle Scholar
- 86.Zheng, G.: Interconnected hollow carbon nanospheres for stable lithium metal anodes. Nat. Nanotechnol. 9, 618–623 (2014)CrossRefPubMedGoogle Scholar
- 87.Ngai, K.S., Ramesh, S., Ramesh, K., et al.: A review of polymer electrolytes: fundamental, approaches and applications. Ionics 22, 1259–1279 (2016)CrossRefGoogle Scholar
- 88.Chen, Y., Tang, Z., Yang, S., et al.: A high-voltage all-solid-state lithium-ion battery with Li–Mn–Ni–O and silicon thin-film electrodes. Mater. Technol. 30, A58–A63 (2015)CrossRefGoogle Scholar
- 89.Schwenzel, J., Thangadurai, V., Weppner, W.: Developments of high-voltage all-solid-state thin-film lithium ion batteries. J. Power Sources 154, 232–238 (2006)CrossRefGoogle Scholar
- 90.Yarmolenko, O.V., Yudina, A.V., Khatmullina, K.G.: Nanocomposite polymer electrolytes for the lithium power sources (a review). Russ. J. Electrochem. 54, 325–343 (2018)CrossRefGoogle Scholar
- 91.Fenton, D.E., Parker, J.M., Wright, P.V.: Complexes of alkali-metal ions with poly(ethylene oxide). Polymer 14, 589 (1973)CrossRefGoogle Scholar
- 92.Long, L., Wang, S., Xiao, M., et al.: Polymer electrolytes for lithium polymer batteries. J. Mater. Chem. A 4, 10038–10069 (2016)CrossRefGoogle Scholar
- 93.Patil, A., Patil, V., Choi, J.W., et al.: Solid electrolytes for rechargeable thin film lithium batteries: a review. J. Nanosci. Nanotechnol. 17, 29–71 (2017)CrossRefGoogle Scholar
- 94.Dong, T.T., Zhang, J.J., Chai, J.C., et al.: Research progress on polycarbonate-based solid-state polymer electrolytes. Acta Polym. Sin. 17, 906–921 (2017)Google Scholar
- 95.Hu, P., Chai, J.C., Duan, Y.L., et al.: Progress in nitrile-based polymer electrolytes for high performance lithium batteries. J. Mater. Chem. A 4, 10070–10083 (2016)CrossRefGoogle Scholar
- 96.Monroe, C., Newman, J.: The impact of elastic deformation on deposition kinetics at lithium/polymer interfaces. J. Electrochem. Soc. 152, A396–A404 (2005)CrossRefGoogle Scholar
- 97.Zhang, D., Yan, H., Zhu, Z., et al.: Electrochemical stability of lithium bis(oxatlato) borate containing solid polymer electrolyte for lithium ion batteries. J. Power Sources 196, 10120–10125 (2011)CrossRefGoogle Scholar
- 98.Choudhury, S., Stalin, S., Deng, Y., et al.: Soft colloidal glasses as solid-state electrolytes. Chem. Mater. 30, 5996–6004 (2018)CrossRefGoogle Scholar
- 99.Seki, S.: Solvent-free 4 V-class all-solid-state lithium-ion polymer secondary batteries. ChemistrySelect 2, 3848–3853 (2017)CrossRefGoogle Scholar
- 100.Lin, Y., Cheng, Y., Li, J., et al.: Biocompatible and biodegradable solid polymer electrolytes for high voltage and high temperature lithium batteries. RSC Adv. 7, 24856–24863 (2017)CrossRefGoogle Scholar
- 101.Li, Q., Imanishi, N., Hirano, A., et al.: Four volts class solid lithium polymer batteries with a composite polymer electrolyte. J. Power Sources 110, 38–45 (2002)CrossRefGoogle Scholar
- 102.Oh, B., Amine, K.: Evaluation of macromonomer-based gel polymer electrolyte for high-power applications. Solid State Ion. 175, 785–788 (2004)CrossRefGoogle Scholar
- 103.Kim, H.S., Lee, C.W., Moon, S.I.: Electrochemical performances of lithium-ion polymer battery using LiNi1/3Co1/3Mn1/3O2 as cathode materials. J. Power Sources 159, 227–232 (2006)CrossRefGoogle Scholar
- 104.Kim, H.S., Kim, S.I., Lee, C.W., et al.: Preparation of lithium-ion polymer battery using LiNi1/3Co1/3Mn1/3O2 as a cathode material and its electrochemical properties. J. Electroceram. 17, 673–677 (2006)CrossRefGoogle Scholar
- 105.Yun, Y.S., Choi, J.A., Kim, D.W.: Lithium polymer batteries assembled with in situ cross-linked gel polymer electrolytes containing ionic liquid. Macromol. Res. 21, 49–54 (2013)CrossRefGoogle Scholar
- 106.Xia, C., Baek, B., Xu, F., et al.: Modification of electrolyte transport within the cathode for high-rate cycle performance of Li-ion battery. J. Solid State Electrochem. 17, 2151–2156 (2013)CrossRefGoogle Scholar
- 107.Park, B., Lee, C.H., Xia, C., et al.: Characterization of gel polymer electrolyte for suppressing deterioration of cathode electrodes of Li ion batteries on high-rate cycling at elevated temperature. Electrochim. Acta 188, 78–84 (2016)CrossRefGoogle Scholar
- 108.Kobayashi, T., Kobayashi, Y., Tabuchi, M., et al.: Oxidation reaction of polyether-based material and its suppression in lithium rechargeable battery using 4 V class cathode, LiNi1/3Mn1/3Co1/3O2. ACS Appl. Mater. Interfaces. 5, 12387–12393 (2013)CrossRefPubMedGoogle Scholar
- 109.Kobayashi, Y., Shono, K., Kobayashi, T., et al.: A long life 4 V class lithium-ion polymer battery with liquid-free polymer electrolyte. J. Power Sources 341, 257–263 (2017)CrossRefGoogle Scholar
- 110.Chen, K., Shen, Y., Jiang, J., et al.: High capacity and rate performance of LiNi0.5Co0.2Mn0.3O2 composite cathode for bulk-type all-solid-state lithium battery. J. Mater. Chem. A 2, 13332–13337 (2014)CrossRefGoogle Scholar
- 111.Shono, K., Kobayashi, T., Tabuchi, M., et al.: Proposal of simple and novel method of capacity fading analysis using pseudo-reference electrode in lithium ion cells: application to solvent-free lithium ion polymer batteries. J. Power Sources 247, 1026–1032 (2014)CrossRefGoogle Scholar
- 112.Chaudoy, V., Ghamouss, F., Luais, E., et al.: Cross-linked polymer electrolytes for Li-based batteries: from solid to gel electrolytes. Ind. Eng. Chem. Res. 55, 9925–9933 (2016)CrossRefGoogle Scholar
- 113.Gao, Y., Zhao, Y., Li, Y.C., et al.: Interfacial chemistry regulation via a skin-grafting strategy enables high-performance lithium-metal batteries. J. Am. Chem. Soc. 139, 15288–15291 (2017)CrossRefPubMedGoogle Scholar
- 114.Zhang, H.P., Zhang, P., Sun, M., et al.: A gelled polymer electrolyte with the blend of PMMA and PVDF of novel stick-like morphology. Z. Phys. Chem. 221, 1039–1047 (2007)CrossRefGoogle Scholar
- 115.Hu, P., Zhao, J.H., Wang, T.S., et al.: A composite gel polymer electrolyte with high voltage cyclability for Ni-rich cathode of lithium-ion battery. Electrochem. Commun. 61, 32–35 (2015)CrossRefGoogle Scholar
- 116.Lee, E.H., Park, J.H., Cho, J.H., et al.: Direct ultraviolet-assisted conformal coating of nanometer-thick poly(tris(2-(acryloyloxy)ethyl) phosphate) gel polymer electrolytes on high-voltage LiNi1/3Co1/3Mn1/3O2 cathodes. J. Power Sources 244, 389–394 (2013)CrossRefGoogle Scholar
- 117.Zeng, X.X., Yin, Y.X., Li, N.W., et al.: Reshaping lithium plating/stripping behavior via bifunctional polymer electrolyte for room-temperature solid Li metal batteries. J. Am. Chem. Soc. 138, 15825–15828 (2016)CrossRefPubMedGoogle Scholar
- 118.Li, X., Qian, K., He, Y.B., et al.: A dual-functional gel-polymer electrolyte for lithium ion batteries with superior rate and safety performances. J. Mater. Chem. A 5, 18888–18895 (2017)CrossRefGoogle Scholar
- 119.Jung, Y.C., Park, M.S., Kim, D.H., et al.: Room-temperature performance of poly(ethylene ether carbonate)-based solid polymer electrolytes for all-solid-state lithium batteries. Sci. Rep. 7, 17482 (2017)CrossRefPubMedPubMedCentralGoogle Scholar
- 120.Tillmann, S.D., Isken, P., Lex-Balducci, A.: Gel polymer electrolyte for lithium-ion batteries comprising cyclic carbonate moieties. J. Power Sources 271, 239–244 (2014)CrossRefGoogle Scholar
- 121.Chai, J., Liu, Z., Zhang, J., et al.: A superior polymer electrolyte with rigid cyclic carbonate backbone for rechargeable lithium ion batteries. ACS Appl. Mater. Interfaces. 9, 17897–17905 (2017)CrossRefPubMedGoogle Scholar
- 122.Damjanovic, D.: Ferroelectric, dielectric and piezoelectric properties of ferroelectric thin films and ceramics. Rep. Prog. Phys. 61, 1267–1324 (1998)CrossRefGoogle Scholar
- 123.Reale, P., Privitera, D., Panero, S., et al.: An investigation on the effect of Li+/Ni2+ cation mixing on electrochemical performance and analysis of the electron conductivity properties of LiCo0.33Mn0.33Ni0.33O2. Solid State Ion. 178, 1390–1397 (2007)CrossRefGoogle Scholar
- 124.Li, Z.H., Zhang, H.P., Zhang, P., et al.: Macroporous nanocomposite polymer electrolyte for lithium-ion batteries. J. Power Sources 184, 562–565 (2008)CrossRefGoogle Scholar
- 125.Yun, Y.S., Kim, J.H., Lee, S.Y., et al.: Cycling performance and thermal stability of lithium polymer cells assembled with ionic liquid-containing gel polymer electrolytes. J. Power Sources 196, 6750–6755 (2011)CrossRefGoogle Scholar
- 126.Hofmann, A., Schulz, M., Hanemann, T.: Gel electrolytes based on ionic liquids for advanced lithium polymer batteries. Electrochim. Acta 89, 823–831 (2013)CrossRefGoogle Scholar
- 127.Ju, S.H., Lee, Y.S., Sun, Y.K., et al.: Unique core–shell structured SiO2(Li+) nanoparticles for high-performance composite polymer electrolytes. J. Mater. Chem. A 26, 395–401 (2013)CrossRefGoogle Scholar
- 128.Manikandan, P., Kousalya, S., Periasamy, P.: Physicochemical characteristics of poly(vinylidene fluoride-hexafluoropropylene)–alumina for mesocarbon microbeads versus LiNi1/3Mn1/3Co1/3O2 Li-ion polymer cells. J. Phys. Chem. Solids 74, 1492–1498 (2013)CrossRefGoogle Scholar
- 129.Yang, C.C., Lian, Z.Y., Lin, S.J., et al.: Preparation and application of PVDF-HFP composite polymer electrolytes in LiNi0.5Co0.2Mn0.3O2 lithium-polymer batteries. Electrochim. Acta 134, 258–265 (2014)CrossRefGoogle Scholar
- 130.Kim, K.W., Kim, H.W., Kim, Y., et al.: Composite gel polymer electrolyte with ceramic particles for LiNi1/3Mn1/3Co1/3O2–Li4Ti5O12 lithium ion batteries. Electrochim. Acta 236, 394–398 (2017)CrossRefGoogle Scholar
- 131.Kong, J.Z., Xu, L.P., Wang, C.L., et al.: Facile coating of conductive poly(vinylidene fluoride-trifluoroethylene) copolymer on Li1.2Mn0.54Ni0.13Co0.13O2 as a high electrochemical performance cathode for Li-ion battery. J. Alloys Compd. 719, 401–410 (2017)CrossRefGoogle Scholar
- 132.Zhang, S.S., Fan, X., Wang, C.: Preventing lithium dendrite-related electrical shorting in rechargeable batteries by coating separator with a Li-killing additive. J. Mater. Chem. A 6, 10755–10760 (2018)CrossRefGoogle Scholar
- 133.Panero, S., Satolli, D., D’Epifano, A., et al.: High voltage lithium polymer cells using a PAN-based composite electrolyte. J. Electrochem. Soc. 149, A414–A417 (2002)CrossRefGoogle Scholar
- 134.Shin, W.K., Cho, J., Kannan, A.G., et al.: Cross-linked composite gel polymer electrolyte using mesoporous methacrylate-functionalized SiO2 nanoparticles for lithium-ion polymer batteries. Sci. Rep. 6, 26332 (2016)CrossRefPubMedPubMedCentralGoogle Scholar
- 135.Park, S.R., Jung, Y.C., Shin, W.K., et al.: Cross-linked fibrous composite separator for high performance lithium-ion batteries with enhanced safety. J. Membr. Sci. 527, 129–136 (2017)CrossRefGoogle Scholar
- 136.Park, J.H., Cho, J.H., Kim, S.B., et al.: A novel ion-conductive protection skin based on polyimide gel polymer electrolyte: application to nanoscale coating layer of high voltage LiNi1/3Co1/3Mn1/3O2 cathode materials for lithium-ion batteries. J. Mater. Chem. 22, 12574–12581 (2012)CrossRefGoogle Scholar
- 137.l’Abee, R., DaRosa, F., Armstrong, M.J., et al.: High temperature stable Li-ion battery separators based on polyetherimides with improved electrolyte compatibility. J. Power Sources 345, 202–211 (2017)CrossRefGoogle Scholar
- 138.Gupta, S.K., Jha, P., Singh, A., et al.: Flexible organic semiconductor thin films. J. Mater. Chem. C 3, 8468–8479 (2015)CrossRefGoogle Scholar
- 139.Zhang, P., Zhang, L., Ren, X., et al.: Preparation and electrochemical properties of LiNi1/3Co1/3Mn1/3O2–PPy composites cathode materials for lithium-ion battery. Synth. Met. 161, 1092–1097 (2011)CrossRefGoogle Scholar
- 140.Xiong, X., Ding, D., Wang, Z., et al.: Surface modification of LiNi0.8Co0.1Mn0.1O2 with conducting polypyrrole. J. Solid State Electrochem. 18, 2619–2624 (2014)CrossRefGoogle Scholar
- 141.Wu, C., Fang, X., Guo, X., et al.: Surface modification of Li1.2Mn0.54Co0.13Ni0.13O2 with conducting polypyrrole. J. Power Sources 231, 44–49 (2013)CrossRefGoogle Scholar
- 142.Wang, D., Li, X., Wang, Z., et al.: Co-modification of LiNi0.5Co0.2Mn0.3O2 cathode materials with zirconium substitution and surface polypyrrole coating: towards superior high voltage electrochemical performances for lithium ion batteries. Electrochim. Acta 196, 101–109 (2016)CrossRefGoogle Scholar
- 143.Gao, Y., Yi, R., Li, Y.C., et al.: General method of manipulating formation, composition, and morphology of solid-electrolyte interphases for stable li-alloy anodes. J. Am. Chem. Soc. 139, 17359–17367 (2017)CrossRefPubMedGoogle Scholar