Next Article in Journal
The Evolution Law of Wettability Degree After Energy Replenishment in Tight Type-II Reservoirs with Different Pore Structures
Previous Article in Journal
Comparative Removal of Hexavalent Chromium from Aqueous Solution Using Plant-Derived and Industrial Zirconia Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multiscale Insights into Inorganic Filler Regulation, Ion Transport Mechanisms, and Characterization Advances in Composite Solid-State Electrolytes

1
Shanghai Key Laboratory of Materials Protection and Advanced Materials in Electric Power, Shanghai Engineering Research Center of Energy-Saving in Heat Exchange Systems, Shanghai University of Electric Power, Shanghai 200090, China
2
Graduate Institute of Applied Science and Technology, National Taiwan University of Science and Technology, Taipei 106335, Taiwan
3
Shanghai Institute of Pollution Control and Ecological Security, Shanghai 200092, China
*
Authors to whom correspondence should be addressed.
Processes 2025, 13(9), 2795; https://doi.org/10.3390/pr13092795
Submission received: 7 August 2025 / Revised: 21 August 2025 / Accepted: 28 August 2025 / Published: 1 September 2025
(This article belongs to the Section Chemical Processes and Systems)

Abstract

All-solid-state lithium batteries (ASSLBs) are emerging as a promising alternative to conventional lithium-ion batteries, offering solutions to challenges related to energy density and safety. Their core advancement relies on breakthroughs in solid-state electrolytes (SEs). SEs can be broadly grouped into two main types: inorganic solid electrolytes (ISEs) and organic solid electrolytes (OSEs). ISEs offer high ionic conductivity (0.1~1 mS cm−1), a lithium-ion transference number close to 1, and excellent thermal stability, but their intrinsic brittleness leads to poor interfacial wettability and processing difficulties, limiting practical applications. In contrast, OSEs exhibit good flexibility and interfacial compatibility but suffer from poor ionic conductivity (10−4~10−2 mS cm−1) due to high crystallinity at room temperature, in addition to poor thermal stability and weak mechanical integrity, making it difficult to match high-voltage cathodes and suppress lithium dendrite growth. Against this backdrop, the stability of the organic–inorganic interface plays a crucial role. However, challenges such as low overall conductivity and unstable interfaces still limit their performance. This review provides a microscopic perspective on lithium-ion transport pathways across the polymer phase, the inorganic filler phase, and their interfacial regions. It categorizes inert fillers and active fillers, analyzing their structure–performance relationships and emphasizing the synergistic effects of filler dimensionality, surface chemistry, and interfacial interactions. In addition, cutting-edge analytical methods such as time-of-flight secondary ion mass spectrometry (TOF-SIMS) and high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) have also been employed and are summarized into their roles for revealing the microstructures and dynamic interfacial behaviors of OICSEs. Finally, future directions are proposed, such as hierarchical pore structure design, surface functionalization, and simulation-guided optimization, aiming to provide theoretical insights and technological strategies for the development of high-performance composite electrolytes for ASSLBs.

1. Introduction

Rechargeable lithium-ion batteries (LIBs) employ volatile and flammable organic liquid electrolytes, which introduce considerable safety concerns, especially in applications such as electric vehicles and large-scale energy storage systems [1,2,3]. A typical lithium-ion battery consists of a cathode, an anode, an electrolyte, and a separator. During charging, lithium ions migrate from the cathode, through the electrolyte, and are then intercalated into the anode material, while electrons flow through the external circuit to maintain charge balance. During discharge, the process is reversed: lithium ions move back to the cathode, and electrons are released to the external circuit to deliver electrical energy. The electrolyte, which facilitates ionic transport while preventing electronic conduction, plays a crucial role in determining the overall electrochemical performance, safety, and longevity of the battery. Furthermore, advanced lithium-ion batteries using liquid electrolytes can now achieve energy densities close to 260 Wh kg−1, approaching their theoretical limit [4,5,6].
Against this backdrop, all-solid-state lithium-metal batteries (ASSLBs), exhibiting a remarkably elevated theoretical energy density and superior intrinsic safety, are recognized as a highly promising technology for the development of future energy storage systems [7,8,9]. As the core component of ASSLBs, solid-state electrolytes (SEs) have recently attracted increasing research interest [10,11], as their physicochemical properties, including interfacial kinetics, system safety, and material lifespan, significantly influence the battery’s overall performance and energy efficiency [12,13].
Solid-state electrolytes (SEs) are generally classified into two categories: inorganic solid-state electrolytes (ISEs) and organic solid-state electrolytes (OSEs). Representative ISEs, such as garnet-type Li7La3Zr2O12 (LLZO) and NASICON-type Li1.3Al0.3Ti1.7(PO4)3 (LATP), showcase excellent ionic transport properties, near-unity Li+ transference efficiency, excellent heat tolerance, and strong mechanical integrity [14]. Nevertheless, their inherent brittleness and high hardness lead to poor wettability at the electrode interfaces, significantly increasing fabrication complexity and introducing uncertainties in practical applications [15,16,17]. In contrast, OSEs, which are typically based on polymer matrices such as polyethylene oxide (PEO), demonstrate excellent elasticity, flexibility, interfacial adhesion, and compatibility [18,19,20]. Nevertheless, their significant degree of crystallinity at room temperature leads to suppressed ionic transport, which fails to meet the demands of high-power applications. Moreover, their insufficient thermodynamic stability (oxidation potential below 4 V) and insufficient mechanical strength significantly compromises the ability to restrain the development of lithium dendrites [21].
To improve the overall performance of organic solid-state electrolytes (OSEs), researchers have developed various strategies, including block or cross-linked copolymerization, plasticizer addition, and the inclusion of inorganic reinforcing agents [22,23,24]. Among the available techniques, OICSEs (organic–inorganic composite solid electrolytes), which take advantage of the synergistic effects of polymeric matrices and inorganic fillers, are widely acknowledged as a viable and efficient method for producing high-performance solid-state electrolytes for ASSLBs [25,26]. Inorganic fillers can broadly be categorized into two types: inert fillers [27,28,29] (such as metal oxides like Al2O3, SiO2, and BaTiO3; carbon materials like graphene oxide (GO); and halloysite nanotubes (HNTs)) and active fillers [30,31,32] (such as sulfide-type Li10GeP2S12 (LGPS), garnet-type LLZO, NASICON-type LATP, and perovskite-type Li0.3La0.6TiO3 (LLTO)). Previous research has demonstrated that the primary functional roles of inorganic fillers include the following mechanisms [33,34,35]: (1) suppressing polymer crystallization and decreasing the polymer’s glass transition point, thus boosting the amorphous phase content and facilitating localized chain mobility; (2) enhancement of lithium-ion transport is achieved via Lewis acid–base interactions involving surface functional groups and lithium salt anions or polymer backbones, where interaction intensity is governed by filler characteristics such as scale, composition, amount, morphology, and functional moieties located on the surface; (3) mitigating anion aggregation at the anode interface enhances the lithium-ion transference number and strengthens the electrochemical stability of OICSEs; and (4) inorganic fillers, when homogeneously dispersed, enhance the mechanical and thermal properties of the electrolyte, consequently elevating the battery system’s overall safety and dependability.
To enhance the electrochemical characteristics of OICSEs, inorganic fillers with diverse dimensional structures, including zero-dimensional particles, one-dimensional nanowires, two-dimensional nanosheets, and three-dimensional frameworks (Figure 1), have been purposefully designed and extensively studied [36,37,38]. The diverse shapes of these fillers contribute to the formation of continuous lithium-ion conduction pathways and establish rapid ion-conducting pathways between the cathode and anode [39]. Compared with inert fillers, active fillers possess intrinsic ionic conductivity, allowing them to directly participate in ion transport processes. Moreover, they can form highly conductive percolation networks with the polymer matrix in OICSEs, which greatly improves the electrochemical efficiency of the entire battery system [40,41].
This review focuses on the critical role of various types of inorganic fillers and their advanced architectures in improving the functional efficiency of OICSEs (Figure 2). First and foremost, key factors such as ionic conductivity, lithium-ion transference number, mechanical properties, and electrochemical stability are analyzed. Next, the effects of filler size, content, morphology, and alignment on ionic conductivity are discussed. From a microscopic perspective, Li+ conduction pathways within OICSEs are comprehensively summarized based on the filler content, category, and system-specific characteristics. In addition, categorizing fillers into inert and active types emphasizes the connection between their structural features and the electrochemical behavior of OICSEs. Finally, the applications and future challenges of advanced characterization techniques, such as TOF-SIMS and HAADF-STEM, are discussed. Through this review of current progress and future directions, we strive to deliver scientific insights and practical guidance to support the creation of high-performance OICSEs.

2. Ion Transport Mechanisms in Composite Solid-State Electrolytes

Current research suggests that ions primarily transport through three types of pathways in organic–inorganic composite electrolytes: ion movement facilitated by the polymer matrix, ion transport through the inorganic active fillers, and transport of ions at the boundary regions between polymer and inorganic fillers.

2.1. Ion Transport in the Polymer Matrix

While ionic species move through the polymer network, the conduction mechanism is consistent with that of the polymer itself. Figure 2 illustrates the conduction mechanisms of polymer electrolytes in two distinct regions [42]. In the amorphous region, polar functional groups inside the polymer provide lone pairs of electrons that weakly coordinate with Li+ ions. The segmental motion of polymer chains enabled by these complexes aids in Li+ migration and ion transport (Figure 2a). In the crystalline region, Li+ ions diffuse via vacancies within helical channels formed by the polymer chains (Figure 2b). However, due to the limited channel diameter, restricted segmental mobility of polymer chains, and coulombic interactions with anions outside the channels, the Li+ transport rate within the crystalline region is considerably lower compared to that in the amorphous region. In crystalline regions, the highly ordered atomic arrangement limits the number of available pathways for Li+ migration, making ion transport relatively slow. By contrast, the amorphous regions possess a more disordered structure with greater free volume and flexible pathways, which facilitates faster Li+ diffusion. Therefore, promoting the formation of the polymer’s amorphous phase is a key strategy for optimizing ion transport performance.

2.2. Ion Transport in Inorganic Active Fillers

When the inorganic phase is composed of active fillers, their intrinsic material properties confer ionic transport capabilities. Due to their ionic conductivity, active fillers have attracted considerable attention in composite electrolytes. For composite electrolyte systems containing active fillers, clarifying ion transport pathways is critical. Currently, the most widely studied active fillers mainly include garnet-type, perovskite-type, NASICON-type, and sulfide-based materials. Zheng et al. [43] utilized isotope labeling alongside high-resolution solid-state lithium nuclear magnetic resonance (NMR) to deeply explore Li+ transport mechanisms. The results indicate that Li+ preferentially diffuses through the LLZO ceramic phase, while the fraction of transport via the interface or the PEO polymer phase is relatively small. Zheng et al. [44] further employed solid-state NMR to compare the changes in 6Li and 7Li signals before and after charge–discharge cycling, elucidating the migration pathways of Li+ in the PEO/LiTFSI/LLZO composite electrolyte system. Their findings show that when the LLZO content reaches 50 wt%, the intensity of the 6Li signal peak corresponding to the LLZO phase significantly increases by 27.2% after cycling, strongly supporting that lithium ions are predominantly transported through the percolation network formed by LLZO, with only a minor portion migrating via the organic phase. This discovery reveals at the molecular scale that the three-dimensional conductive network formed by LLZO fillers dominates the ion transport mechanism in composite electrolytes. Zhu et al. [45] found that their developed PEO/LiTFSI/Li0.3La0.6TiO3 (LLTO) composite electrolyte system exhibits significantly enhanced ionic conductivity, which can be attributed to a dual mechanism: on one hand, the 15 wt% LLTO filler effectively suppresses crystallization in the PEO matrix; on the other hand, abundant vacancy defects in the LLTO crystal structure provide special channels for Li+ migration, where lithium ions migrate via a vacancy hopping mechanism along the LLTO lattice. Introducing ion-conductive fillers like LLTO into the polymer matrix has been shown to both regulate polymer microstructure, thereby improving ion transport efficiency, and enable direct ion conduction via vacancy-mediated pathways in the inorganic phase. Collectively, these results confirm that inorganic active fillers can establish multidimensional ion transport networks within polymer composite electrolytes. Additionally, Liu et al. [46] developed a composite electrolyte system that breaks traditional lithium salt dependence, achieving effective ion conduction solely by compositing polymer PEGDA with the active filler LATP. Similarly, Zhang et al. [47] made important advances in lithium salt-free electrolyte design by directly compositing nano-sized LLZTO particles with PEO, yielding a composite electrolyte exhibiting ionic conductivity measured at 2.1 × 10−4 S·cm−1 at 30 °C, significantly exceeding that of pure PEO matrix. This breakthrough not only confirms that inorganic active fillers serve as three-dimensional ion transport channels but more importantly reveals their unique role as lithium-ion sources, where the fillers’ own Li+ release provides mobile charge carriers directly to the composite system, thereby overcoming the traditional electrolyte’s reliance on lithium salt additives.

2.3. Ion Transport at Polymer/Inorganic Filler Boundaries

Besides the polymer and inorganic active fillers, a considerable number of studies have demonstrated the existence of a third ion transport pathway within organic–inorganic composite electrolytes, namely, the fast ion transport interface at the organic–inorganic phase boundary. The formation mechanism of this special interfacial region is primarily driven by the topological disruption effect of inorganic fillers: when inorganic fillers are dispersed throughout the polymer matrix, strong interactions between their surfaces and polymer chain segments disrupt the original molecular ordering, forming an amorphous network with high free volume in the interfacial region. Specifically, taking the LLZO/PEO system as an example, the rigid surface of zirconia-based fillers in contact with flexible PEO chains produces a disordered transition layer approximately 5–10 nm thick, where the chain segment relaxation time is shortened by 2–3 orders of magnitude compared to the bulk polymer. This structural rearrangement significantly reduces the activation energy barrier for ion migration, resulting in an interfacial ionic conductivity 3–5 times higher than that of the bulk polymer, thereby constituting a third fast ion transport channel distinct from the intrinsic conduction in either organic or inorganic phases. This finding breaks through the traditional two-phase conduction model framework and provides a new dimension for designing multi-scale ion transport networks. Lin et al. [48] constructed a PEO/SiO2 composite electrolyte system with strong interfacial coupling via an in situ hydrolysis technique. Using an in situ synthesis strategy, silica nanoparticles with a diameter of 12 nm were prepared, whose surface-abundant silanol groups formed hydrogen bonding networks with the ether oxygen groups of PEO molecules (binding energy, approximately 25 kJ mol−1). This strong interaction reduced the crystallinity of PEO from 65% in the pure phase to 38%, while the free volume fraction increased by 1.8 times. Molecular dynamics simulations revealed that within an interfacial region of approximately 3 nm, the PEO chain segment relaxation time was shortened to 1/20 of that in the bulk phase, significantly promoting lithium-ion migration via the cooperative motion of polymer chains. As a result, the composite electrolyte exhibited an ionic conductivity of 0.56 mS cm−1 at 60 °C, representing an increase of two orders of magnitude over the pure PEO system. This work elucidates at the molecular interaction level the microscopic mechanism by which inert fillers regulate polymer dynamics through interfacial engineering. Wan et al. [49] incorporated inorganic active LLZO fillers in the form of nanowires into a PEO/LiTFSI system and found that the high specific surface area of LLZO nanowires (210 m2 g−1) significantly decreased the crystallinity of the polymer electrolyte from 72% in the pure phase to 41%, thereby increasing the ionic conductivity of the composite electrolyte at 45 °C to 0.47 mS cm−1, nearly an order of magnitude higher than the pure PEO system. The study also revealed a secondary interfacial enhancement mechanism: La-O basic sites enriched on the LLZO surface (isoelectric point pH = 9.2) strongly coordinate with TFSI- anions from LiTFSI (binding energy ΔG = −18.6 kJ mol−1), increasing the lithium salt dissociation degree by 53% and resulting in a free Li+ concentration at the interface region 2.8 times higher than that in the bulk phase. This work demonstrates multidimensional optimization of ion transport performance in composite electrolytes through the synergistic control of morphology and interfacial chemistry. Xu et al. [50] investigated the PEO/Li0.38Sr0.44Ta0.75Zr0.25O3 composite electrolyte system and revealed the key regulatory role of interfacial chemistry on lithium-ion transport. X-ray photoelectron spectroscopy (XPS) and density functional theory (DFT) calculations confirmed that lone pair electrons of fluorine in the TFSI- lithium salt (binding energy 532.1 eV) form strong Lewis acid–base interactions with empty d orbitals of tantalum atoms in the filler (Ta5+ d orbital occupancy ≈ 1.2), which specifically lowers the conformational entropy of TFSI- by 37% and reduces the lithium salt dissociation barrier by 0.45 eV. Extended X-ray Absorption Fine Structure (EXAFS) analysis showed that the Li0-O coordination number at the interface decreased from 4.2 in the bulk to 3.5, indicating partial decoupling of the Li+ solvation shell and increasing the free Li+ concentration to 1.8 × 1021 cm−3. Such molecular-scale interfacial engineering resulted in a Li+ transference number of 0.62 at 50 °C, exceeding the pure PEO system by 42%, confirming the directional regulation ability of inorganic filler surface chemistry on interfacial ion transport kinetics. Hu et al. [51] innovatively employed a synchronized electrospinning technique to fabricate composite electrolytes with a three-dimensional continuous ion transport network. By precisely controlling the electric field parameters (18 kV voltage, 15 cm collector distance), molecular-level interfacial coupling between the polymer matrix and ZrO2 nanoparticles (average particle size 25 nm) was achieved during fiber formation, resulting in advanced electrolyte materials exhibiting a bulk conductivity of 1.16 mS cm−1 and an areal conductivity of 464 mS cm−2. Mechanistic studies revealed that acidic sites enriched on the ZrO2 surface (isoelectric point pH = 4.2) form strong Lewis acid–base interactions with lone pair electrons of oxygen atoms in ClO4 anions (binding energy ΔG = −22.3 kJ mol−1), improving the dissociation level of lithium salt by 68% and elevating the free Li+ concentration at the interface to 3.4 × 1021 cm−3. TOF-SIMS three-dimensional imaging demonstrated that the Li+ diffusion coefficient at the ZrO2/PEO interface (2.7 × 10−10 cm2 s−1) was two orders of magnitude higher than that in the bulk phase, confirming the decisive role of interfacial engineering in constructing fast ion channels. This research provides a novel interface design paradigm for overcoming performance bottlenecks in traditional electrolyte materials. Finally, a space charge layer is generally established at the boundary of inorganic fillers and the organic phase, further promoting interfacial ion transport. Li et al. [52] directly visualized a space charge layer roughly 3 nm in thickness at the interface of a PEO/Ga-LLZO composite electrolyte using aberration-corrected transmission electron microscopy (AC-TEM). Phase-field dynamics simulations indicated that lattice distortion of LLZO caused by Ga3+ doping (lattice expansion rate 1.8%) increased the oxygen vacancy concentration on the filler surface to 1.2 × 1020 cm−3, creating a potential gradient of approximately 0.35 V nm−1 at the interface. This built-in electric field reduced the migration activation energy of Li+ within the space charge layer from 0.45 eV in the bulk to 0.28 eV. Monte Carlo simulations based on a three-dimensional random resistor network model (with more than 106 nodes) showed that when the filler volume fraction exceeds 23%, the percolation network formed by the space charge layer contributes 62% of the total conductivity. The system achieved an ionic conductivity of 0.14 mS cm−1 at 30 °C, roughly three orders of magnitude above the conductivity of pure PEO. This interfacial engineering strategy, by regulating the chemical potential gradient (Δμ ≈ 0.12 eV) of the space charge layer, increases the Li+ diffusion coefficient at the interface to 3.8 × 10−10 cm2 s−1, two orders of magnitude greater than that in the bulk polymer. These breakthroughs confirm that constructing optimized space charge layers not only enhances conductive properties of composite electrolytes under traditional percolation thresholds but also provides theoretical guidance for designing high-ion-flux electrolytes for all-solid-state batteries.

3. Key Inorganic Fillers and Advanced Structures in Organic–Inorganic Composite Solid Electrolytes

Organic–inorganic composite solid electrolytes (OICSEs) consist of a polymer matrix, lithium salt, and inorganic fillers. In 1973, Wright et al. [53] initially suggested that lithium ions could be conducted by mixing alkali metal salts with polyethylene oxide (PEO). As of now, common polymer matrices often comprise PEO [54], polyvinylidene fluoride and hexafluoropropylene [55], polyvinylidene fluoride [56], polyethylene glycol diacrylate [46], polymethyl methacrylate [57], polyvinyl carbonate [58], tetraethylene glycol dimethacrylate monomer [59], and polymer polystyrene, among others. Generally, these polymers display semi-crystalline behavior at room temperature, restricting segmental motion and resulting in low ionic conductivities [60]. Ionic conductivity is significantly enhanced as polymers enter the amorphous phase upon exceeding the glass transition temperature. In general, lithium salts are grouped into inorganic and organic types. Inorganic lithium salts include lithium perchlorate (LiClO4), lithium tetrafluoroborate (LiBF4), lithium hexafluoroarsenate (LiAsF6) [61], and lithium hexafluorophosphate (LiPF6). Organic lithium salts include lithium bis(oxalato)borate (LBOB), lithium difluoro(oxalato)borate (LDFOB), lithium bis(fluorosulfonyl)imide (LFSI), and lithium bis(trifluoromethanesulfonyl)imide (LTFSI), the latter exhibiting high solubility in polymers and rapid formation of stable solid electrolyte interphase (SEI) films. Inorganic fillers are categorized into inert and active fillers based on their lithium-ion conductivity. Inert fillers, which do not participate in conduction, include ZnO, TiO2, SiO2, ZrO2, MgO, Al2O3, Y3O3 LiAlO2, and BaTiO3 [62]. Their functional mechanisms involve inhibiting polymer crystallization, promoting lithium-ion migration through Lewis acid–base interactions between surface functional groups and lithium salt anions or the polymer matrix, increasing lithium-ion transference number and suppressing anion accumulation on the anode side, as well as improving electrolyte mechanical strength and thermal stability. Active fillers include garnet-type, sulfide-type, NASICON-type, LISICON-type, and Li3N. Due to their inherent ionic conductivity, active fillers directly participate in ion transport and construct fast ion-conductive percolation pathways with the polymer matrix in OICSEs. The shapes of inorganic fillers vary widely, encompassing 0D nanoparticles, 1D nanofibers, 2D nanosheets, and 3D frameworks. Fillers of different morphologies form long-range percolation networks through ordered arrangements, promoting lithium-ion conduction and enhancing diffusion rates, thereby creating rapid ion transport channels.

3.1. Inert Fillers

3.1.1. Zero-Dimensional Inert Fillers

Particles classified as zero-dimensional (0D) inert materials typically measure from several nanometers to a few micrometers. Adding these fillers to polymer electrolytes that contain lithium salts can improve structural strength, ionic conductivity, and electrochemical performance. This enhancement in performance is primarily due to the suppression of polymer crystallization by inert fillers, which enhances the segmental mobility of polymer chains. Additionally, surface interactions involving Lewis acid–base chemistry on nanoparticles and polyethylene oxide (PEO) chain segments also facilitate lithium salt dissociation, thus attracting widespread attention. Croce et al. [63] confirmed that in PEO-based organic–inorganic composite solid electrolyte (OICS) systems, the enhancement of electrochemical performance is related to the participation of surface hydroxyl groups (-OH) on Al2O3 within the polymer matrix via hydrogen bonding facilitated by anions during the solvation process. This mechanism reduces lithium salt association and reduces the likelihood of salt formation, thereby enhancing specific interactions between fillers and polymer chains (Figure 3a). Directed modification of inert fillers via surface engineering can significantly improve ionic transport in composite electrolytes. Specifically, regulating the density of acidic sites on filler surfaces (e.g., increasing Al2O3 surface hydroxyl concentration to 8.5 OH nm−2) or chemical functionalization (e.g., amino-functionalized SiO2) yields a dual effect. First, intense Lewis acid–base interactions occurring at the interface between filler and polymer (e.g., coordination bonding between Al3+ and PEO ether oxygens with binding energies of 18–25 kJ mol−1) disrupt the regular alignment of polymer segments, reducing crystallinity by 30–50% and lowering segmental motion activation energy from 1.2 eV to 0.6 eV. Second, acidic sites on filler surfaces (e.g., isoelectric point pH = 4.2 for ZrO2) form specific coordination with lithium salt anions (e.g., fluorine atoms in TFSI-), lowering dissociation energy by 0.3–0.5 eV and increasing free Li+ concentration in the interface region to 2–3 times that of the bulk phase. For instance, TiO2 nanoparticles modified by phosphorylation can increase LiTFSI dissociation from 52% to 78%, corresponding to an increase in ionic conductivity by two orders of magnitude. This precise interfacial chemical engineering provides important theoretical guidance for designing high-performance composite electrolytes.
Xue et al. [64] introduced urea–pyrimidinone functionalized SiO2 into a polymer matrix containing UPy units, constructing a dynamic cross-linked network based on quadruple hydrogen bonds (Figure 3b). The hydrogen bonding between SiO2-UPy and PEG-UPy (bond energy ≈ 60 kJ mol−1) improved filler dispersion threefold and increased free volume fraction to 0.28. The dynamic reversible hydrogen bond network enabled the material to retain 92% of its ionic conductivity (8.0 × 10−5 S·cm−1 at 30 °C) after five break–heal cycles. Yang et al. [65] employed in situ self-assembly to fabricate a PEO@SiO2 three-dimensional interpenetrating network (Figure 3c). Monodisperse SiO2 nanoparticles (~15 nm) strongly interacted via Lewis acid–base coordination (Al3+-O bonds, binding energy 18 kJ mol−1) and weak hydrogen bonding, reducing PEO crystallinity from 65% to 38% and forming an approximately 5 nm thick amorphous interfacial region. This system achieved an ionic conductivity of 1.1 × 10−4 S·cm−1 at 45 °C, and Li||Li symmetric cells demonstrated stable cycling for 1000 h at 0.2 mA cm−2. Park et al. [66] incorporated extensively mesoporous SiO2 nanoparticles into a poly(propylene carbonate) (PPC) matrix (Figure 3d). The strong Lewis acidic sites on the internal surfaces anchored TFSI- anions, resulting in a lithium-ion transport number measured at 0.86. Molecular dynamics simulations indicated that Li+ diffusion coefficients at the mesoporous silica nanoparticle interface reached 3.8 × 10−10 cm2 s−1, five times higher than the bulk matrix. The assembled NCM622 full cells maintained 91.2% capacity retention after 500 cycles at 1C. These studies demonstrate multiscale synergistic optimization from molecular-level interactions to macroscopic ion transport networks by regulating filler surface chemistry (acid site density and functional group modification) and topology (3D networks and mesoporous channels), providing crucial technical routes for developing high-safety solid-state batteries.

3.1.2. One-Dimensional Inert Fillers

In organic–inorganic composite solid electrolyte (OICSE) systems, the introduction of one-dimensional (1D) inert fillers significantly enhances ionic transport performance through multiple mechanisms. Compared with zero-dimensional fillers, 1D nanowires/nanorods possess a unique continuous topology that can surpass traditional percolation threshold limits, forming long-range ion percolation networks. For example, Cui et al. [67] developed an yttria-stabilized zirconia (YSZ) nanowire system, where 7 mol% Y-doped ZrO2 nanowires feature positively charged oxygen vacancies on their surface (density, approximately 1.2 × 1021 cm−3). These sites act as Lewis acids binding to TFSI- anions, resulting in a PAN-based composite electrolyte exhibiting an ionic conductivity of 0.11 mS cm−1 at 25 °C, which is 3.6 times higher than that of a comparable nanoparticle-based system (Figure 4a). Tao et al. [68] incorporated Mg2B2O5 nanowires (diameter ~ 15 nm) into a PEO matrix, where surface Mg2+ ions serve as Lewis acid centers strongly interacting with TFSI- (binding energy ΔG = −14.7 kJ mol−1), enhancing the Li+-TFSI- dissociation degree to 68%. This composite achieved a high ionic conductivity of mS cm−1 at 30 °C (Figure 4b). Hua et al. [69] prepared a TiO2 nanorod/PPC composite system, which, through the synergistic effects of the nanorods’ porous surface structure (pore size ~ 3 nm) and Lewis acid–base sites, attained an ionic conductivity of 0.15 mS cm−1 and a tensile strength of 27 MPa at room temperature. The Li||OICSE||SS asymmetric cell using this electrolyte demonstrated a stable electrochemical window of up to 4.6 V.
Surface chemical modification strategies have further optimized the interfacial transport properties of 1D fillers. Li et al. [70] employed toluene-2,4-diisocyanate (TDI) to modify TiO2 nanowires, constructing a cross-linked network (Figure 4c). The isocyanate groups of TDI form covalent bonds with the ether oxygen groups of PEO chains, leading to a composite electrolyte exhibiting an ionic conductivity of 0.1 mS cm−1 at 25 °C. Zhao et al. [71] coated TiO2 nanofibers with polydopamine (PDA) (Figure 4d); the strong lithium affinity of the PDA layer (contact angle < 15°) creates a rapid conduction layer approximately 2 nm thick at the interface between the filler and polymer. The Li+ diffusion coefficient in the interfacial region reaches 4.8 × 10−10 cm2 s−1, achieving a boosted ionic conductivity of 0.44 mS cm−1 at 55 °C. These studies indicate that 1D fillers suppress polymer chain recrystallization through geometric confinement effects (reducing crystallinity by 20–40%), and their surface chemical features (oxygen vacancies and Lewis acid sites) synergize with their topological structure (aspect ratio > 50) to establish multi-scale ion transport channels. This provides new insights for developing composite electrolytes with high ionic flux.

3.1.3. Two-Dimensional and Three-Dimensional Inert Fillers

In research focused on OICSEs, the incorporation of 2D and 3D inert fillers has significantly optimized ionic transport performance through structural innovations. Two-dimensional materials such as graphene oxide (GO), with abundant surface oxygen-containing functional groups (-COOH and -OH), form strong interactions with the polymeric framework. Xu et al. [72] reported that incorporating 1 wt% GO into a PEO matrix led to an ionic conductivity of 1.54 × 10−2 mS cm−1 at 25 °C and an increased Li+ transference number of 0.42. The assembled LiFePO4 full cells exhibited a capacity retention of 91% after 100 cycles at a 0.5 C rate (Figure 5a). Zhang et al. [73] developed a lithium montmorillonite (LiMNT)/polycarbonate (PEC) composite system that, through a layered structure enabling charge-selective adsorption (Figure 5b), achieved a Li+ transference number of 0.83 at 30 °C. The ordered interlayer arrangement shortened the Li+ transport path to 1.2 nm. MXene-mSiO2 composite nanosheets [74], combining a mesoporous structure (pore size ≈ 5 nm) and surface polar groups (-OH, -F), enabled the PEO-based electrolyte to reach a high ionic conductivity of 0.46 mS cm−1 at 25 °C. The assembled LiFePO4 full cells demonstrated a capacity retention of 141.8 mAh g−1 after 250 cycles, representing a 135% improvement compared to the pure polymer system.
Three-dimensional (3D) scaffold structures overcome the limitations of traditional fillers by constructing continuous ion conduction channels. Zhang et al. [75] employed vertically aligned Al2O3 scaffolds (Figure 5c) to form high aspect-ratio interfaces, with predicted Li+ ionic conductivity along the ceramic/polymer interface exceeding 1 mS cm−1 at 0 °C. The measured ionic conductivity at room temperature reached 0.58 mS cm−1, representing four orders of magnitude improvement over conventional nanowire systems. Cui et al. [76] developed a 3D SiO2 aerogel network (Figure 5d) via a sol–gel method, producing an interpenetrating structure with a modulus of 0.43 GPa and ionic conductivity surpassing 6 × 10−4 S·cm−1 at 30 °C. The continuous adsorption interface reduced anion migration resistance by 78%. Zhang et al. [77] embedded ionic liquids into a glass fiber cloth-based 3D network. The Li||Li symmetric cell maintained stable polarization during 2000 h of cycling. Studies show that the 3D structure reduces polymer crystallinity to 18% through geometric confinement, and its interconnected pores (porosity > 80%) increase Li+ diffusion coefficients in the interfacial region to 1.2 × 10−9 cm2 s−1, three times that of 2D materials. Despite the remarkable advantages of 3D systems, complex fabrication methods, such as directional freezing and electrospinning and challenges in interface engineering remain bottlenecks for practical application. Future research should focus on simplifying fabrication processes (e.g., solution blow spinning technology [78]), optimizing material combinations (e.g., MXene/BN hybrids), and improving electrode compatibility to achieve scalable applications of high-safety solid-state batteries.
In the development of organic–inorganic composite solid electrolytes (OICSEs), inert fillers of different dimensionalities exhibit unique advantages and limitations due to their structural characteristics (Table 1). Zero-dimensional (0D) inert materials, such as nanoparticles, serve as fundamental modifiers owing to their high mechanical strength (modulus > 1 GPa) and chemical stability (thermal decomposition temperature > 400 °C). However, their low specific surface area (<50 m2 g−1) leads to poor interfacial contact, resulting in ionic conductivities typically below 10−3 mS cm−1. One-dimensional (1D) nanowires or nanotubes form oriented conduction pathways due to their high aspect ratio (>50), effectively suppress polymer crystallinity (reducing crystallinity by 20–40%), and enhance interfacial contact (interface thickness ~3 nm). Yet, their complex fabrication methods (such as templating and electrospinning) limit large-scale application. Two-dimensional (2D) nanosheets possess ultrahigh specific surface areas and abundant surface functional groups, which significantly enhance lithium salt dissociation through Lewis acid–base interactions, increasing dissociation degrees by 50%. However, their interlayer stacking effect causes low mechanical strength (tensile strength < 10 MPa). Three-dimensional (3D) scaffold materials construct continuous conductive interfaces via interconnected pores, enabling ionic conductivities exceeding 10−1 mS cm−1. Nonetheless, their fabrication relies on complex techniques, and interface engineering remains challenging. Future research should focus on the synergistic design of multidimensional materials, low-cost fabrication technologies, and cross-scale interface optimization to balance conductivity, mechanical strength, and processing feasibility.

3.2. Active Fillers

In solid-state electrolyte systems, active fillers are key components for enhancing composite electrolyte performance attributable to their built-in ionic conductivity and electrochemical activity. Compared to inert fillers, active materials optimize system performance through a dual mechanism: first, the active phase’s own three-dimensional ion channels (e.g., the cubic phase framework pores of LLZO, approximately 0.6 nm) provide efficient Li+ transport pathways; second, the space charge layer formed at the filler/polymer interface (2~5 nm thick) reduces the Li+ migration activation energy via an internal electric field (potential gradient ≈ 0.3 V nm−1). For example, sulfide fillers (such as Li3PS4), with their superionic conductor properties, enable composite electrolytes to achieve high conductivities exceeding 10−1 mS cm−1 at 25 °C.
However, the introduction of active fillers may cause intensified electrode polarization (interface impedance increasing by 30~50%) and capacity fading (capacity retention below 80% after 100 cycles). The root causes are (1) chemical compatibility issues between the filler and electrode (e.g., interface reactions between garnet-type LLZO and lithium metal forming a Li2CO3 passivation layer); (2) transport kinetic imbalance caused by Li+ concentration gradients at multiphase interfaces. To address these challenges, researchers have proposed multi-scale optimization strategies: at the material level, element doping (e.g., Ta-doping in LLZO to enhance structural stability) and surface coating (e.g., Li3PO4 modification of Li7La3Zr2O12 to suppress interfacial side reactions); at the structural design level, constructing three-dimensional interpenetrating networks to reduce percolation thresholds (<15 vol%) and suppress grain boundary impedance growth through topological constraints. Recent studies show that perovskite-type Li0.33La0.55TiO3 (LLTO) fillers, due to their layered vacancy channels (vacancy concentration ≈ 1021 cm−3) and synergistic transport with the polymer amorphous phase, can increase the Li+ transference number of composite electrolytes up to 0.82, with assembled Li||Li symmetric cells cycling stably for over 2000 h at 0.5 mA cm−2. Future research should focus on precise regulation of filler/matrix interfacial dynamics and cross-scale optimization of transport networks to balance high ion flux and long-term cycling stability.

3.2.1. Polymer Matrices Incorporated with Sulfide-Type Materials

In OICSEs, sulfide active fillers have become a research hotspot due to their high ionic conductivity. However, their poor interfacial stability with lithium metal and sensitivity to air still require optimization through composite strategies. Xu et al. [79] incorporated Li10GeP2S12 (LGPS) into a PEO matrix; with a 1 wt% addition, the electrolyte achieved ionic conductivities of 1.17 × 10−2 mS cm−1 and 1.21 mS cm−1 at 25 °C and 80 °C, respectively. LGPS suppressed PEO crystallinity (reducing crystallinity by 35%) and weakened the Li+-ether oxygen interactions, enabling the LiFePO4||Li battery to retain 92.5% capacity after 50 cycles at 60 °C. To promote more effective filler spreading, Xu et al. [80] proposed an in situ synthesis method for Li3PS4 (Figure 6a): compared with mechanical mixing, the in situ generated nanoparticles (≈20 nm in size) showed a 2.3-fold improvement in uniformity within the PEO matrix. With a 2 vol% addition, the electrolyte achieved an ionic conductivity of 0.80 mS cm−1 at 60 °C, and the assembled LiFePO4 battery retained 81% capacity after 325 cycles.
Addressing the air stability issue of sulfides, Wang et al. [81] developed a Li7PS6/PVDF-HFP composite system, where the hydrophobic nature of PVDF-HFP caused only an 8% decrease in ionic conductivity after exposure to 50% humidity for 24 h. The electrolyte exhibited a room temperature ionic conductivity of 0.11 mS cm−1, and the Li||Li symmetric cell demonstrated stable cycling for 1000 h at 0.2 mA cm−2 (Figure 6b). For flexible design, Liu et al. [82] fabricated an ultrathin Li6PS5Cl/P(VDF-TrFE) electrolyte (<20 μm thick) via electrospinning and hot-pressing; its three-dimensional interpenetrating network structure achieved an ionic conductivity of 1.2 mS cm−1. The LiNi0.8Co0.1Mn0.1O2 full cell maintained 71% capacity following 20,000 cycles under 1.0 mA cm−2. Zhang et al. [83] designed a Li6PS5Cl/PEO film (65 μm thick), which under high loading (4.46 mAh cm−2) enabled LiNi0.7Co0.2Mn0.1O2 batteries to retain 74% capacity after 1000 cycles at 60 °C, with an average Coulombic efficiency of 99.85%.
Figure 6. (a) Schematic diagram representing the in situ preparation pathway of PEO/Li3PS4 hybrid polymer electrolytes. Reproduced with permission from Wang, H. ACS. Appl. Mater. Interfaces; published by American Chemical Society, 2020 [80]. (b) Schematic illustration of SCE (LPS/PVDF-HFP/LiTFSI). Reproduced with permission from Shen, Y. Adv. Energy Mater.; published by Wiley, 2022 [81]. (c) HRTEM of Li2O crystals distributed in the interface layer. (d) Morphology of deposited Li on a Cu grid in a S-CPE and corresponding element mapping. Reproduced with permission from Lee, S. M. J. Power Sources; published by Elsevier, 2015 [84].
Figure 6. (a) Schematic diagram representing the in situ preparation pathway of PEO/Li3PS4 hybrid polymer electrolytes. Reproduced with permission from Wang, H. ACS. Appl. Mater. Interfaces; published by American Chemical Society, 2020 [80]. (b) Schematic illustration of SCE (LPS/PVDF-HFP/LiTFSI). Reproduced with permission from Shen, Y. Adv. Energy Mater.; published by Wiley, 2022 [81]. (c) HRTEM of Li2O crystals distributed in the interface layer. (d) Morphology of deposited Li on a Cu grid in a S-CPE and corresponding element mapping. Reproduced with permission from Lee, S. M. J. Power Sources; published by Elsevier, 2015 [84].
Processes 13 02795 g006
In terms of interface engineering, Su et al. [84] revealed the unique role of sulfide fillers through Cryo-TEM: at the Li/S-CPE interface, LSPSCl promotes the decomposition of TFSI- to generate Li2O nanocrystals (particle size ≈ 5 nm) and an amorphous LiF/Li2S layer (Figure 6c). This composite interface reduces the lithium deposition overpotential to 18 mV, effectively suppressing dendrite growth. EDS elemental mapping (Figure 6d) shows a gradient distribution of O, F, and S elements in the S-CPE, confirming directional regulation of interfacial chemistry. The assembled solid-state battery retained 97.8% capacity after 100 cycles at 0.1 A g−1.
Although sulfide-based composite electrolytes show great potential, their large-scale application still faces challenges: (1) interfacial side reactions between sulfides and polar polymers need to be further suppressed through surface coatings (e.g., Li3PO4 layers) or doping (e.g., Ge substitution); (2) high sulfide content (>30 vol%) often leads to mechanical performance degradation, requiring three-dimensional scaffolds (such as electrospun nanofiber networks) to balance ionic conductivity and mechanical strength; and (3) air stability and compatibility with wet processing still need improvement, for example, by using hydrophobic polymers (like polytetrafluoroethylene) coatings or developing in situ solid-state reaction synthesis methods. Future research should focus on multiscale interface design and high-throughput screening to achieve safe, long-life solid-state battery systems.

3.2.2. Polymer Matrices Incorporated with Garnet-Type Materials

Garnet-type solid-state electrolyte materials (such as Li7La3Zr2O12 and its derivatives) exhibit significant advantages in composite electrolytes due to their high ionic conductivity, wide electrochemical window, and excellent dendrite suppression capability. Lee et al. [85] found that a PEO-based composite electrolyte containing 52.5 wt% cubic-phase LLZO achieved an ionic conductivity of 0.44 mS cm−1 at 55 °C, which is four orders of magnitude higher than a system with the same content of Al2O3 inert filler. This improvement is attributed to the synergistic conduction mechanism between the intrinsic ion channels of the active LLZO filler and the polymer amorphous phase. He et al. [49] constructed a three-dimensional continuous transport network by introducing LLZO nanowires (Figure 7a); their high aspect ratio (>100) enabled an ionic conductivity of 0.24 mS cm−1 at 25 °C. The Li||Li symmetric cell cycled stably for 1000 h at 60 °C without dendrite penetration, and the assembled LiFePO4 full cell retained a capacity of 158.8 mAh g−1 following 70 cycles at 0.5 C. To address the issue of Li2CO3 passivation layer formation on LLZO surfaces, Huang et al. [86] coated LLZTO nanoparticles with polydopamine (PDA). The amphiphilic nature of PDA enabled uniform dispersion of 80 wt% filler in the PEO matrix; after modification, the electrolyte’s ionic conductivity increased from 6.3 × 10−2 to 0.11 mS cm−1 at 30 °C, a 75% improvement compared to the unmodified system. Li et al. [87] developed molecular brush-modified MB-LLZTO fillers (Figure 7b), whose grafted PMImCl groups electrostatically anchored TFSI- anions, resulting in a composite electrolyte with 15 wt% filler content, achieving an ionic conductivity of 0.31 mS cm−1 at 50 °C. Lithium–sulfur batteries exhibited a high initial capacity of 1280 mAh g−1 at low temperature. Yu et al. [88] constructed a polymer–ceramic “bridging” interface (Figure 7c) utilizing chemical and hydrogen bonds to form ultrafast Li+ transport channels, pushing room temperature ionic conductivity beyond 3.1 mS cm−1. Li||Li symmetric cells cycled stably for 1000 h at 0.1 mA cm−2, and LiFePO4 full cells retained 92% capacity after 100 cycles. Regarding 3D structural design, Hu et al. [89] used wood as a template to fabricate low-tortuosity garnet frameworks (Figure 7d), whose aligned channels yielded a room temperature ionic conductivity of 0.18 mS cm−1, approaching the intrinsic bulk conductivity of LLZO. Notably, the compatibility with different polymer matrices significantly affects performance: PVDF-based LLZO composite electrolytes [90] demonstrated a 5.4 V electrochemical window and a Li+ transference number of 0.70, while PEGDA-based systems [91] formed 3D interpenetrating networks via photopolymerization, achieving room temperature ionic conductivity of 0.31 mS cm−1. These studies indicate that morphology control (nanowires and 3D frameworks), surface engineering (molecular brushes and PDA coating), and interfacial bonding design effectively overcome bottlenecks such as poor garnet filler dispersion and high interfacial impedance, providing crucial technical pathways for developing high-safety solid-state batteries.

3.2.3. Polymer Matrices Incorporating NASICON-Type Materials

NASICON-type fast ion conductors have attracted significant attention in composite electrolyte systems due to their high room-temperature ionic conductivity and large electrochemical operating window (~5 V), but their interfacial reactivity with lithium metal remains an urgent issue to be addressed. Wang et al. [92] prepared LATP-PEO hybrid electrolytes (EO/Li = 16) via solution casting, achieving an ionic conductivity of 2.6 × 10−3 mS cm−1 at 25 °C, while a PEO-LiClO4-LAGP system with 15 wt% LAGP (EO/Li = 8) showed enhanced conductivity up to 8.0 × 10−3 mS cm−1. Xin et al. [93] fabricated three-dimensional silane-modified LATP/PVDF composite electrolytes by electrospinning (Figure 8a), where the grafted polysiloxane -NH3+ groups imparted a positive charge to LATP and fully exposed its Lewis acid sites; enhanced anion adsorption via electrostatic interactions led to ionic conductivity of 1.06 mS cm−1 and Li+ transference number of 0.82 at 25 °C, exhibiting 15.3 MPa tensile strength and an expanded electrochemical window of up to 4.86 V. Fan et al. [94] innovatively used NaCl powder as a sacrificial template to prepare a 3D porous LATP framework (Figure 8b); porosity was controlled by tuning template dissolution (Figure 8c). After compositing with PEO, ionic conductivity reached 0.75 mS cm−1 at 60 °C, and Li||Li symmetric cells cycled stably at 0.2 mA cm−2 for 1000 h. To improve filler dispersion, Xing et al. [95] PMMA-coated LATP engineered within a PVDF polymer matrix (Figure 8d); the interaction at the molecular level between PMMA and PVDF promoted the formation of a continuous 3D LATP network, and the Li+ complexation capability of PMMA led to an ultrahigh ionic conductivity of 1.23 mS cm−1 and a Li+ transference number of 0.85 at room temperature. In terms of interface engineering, Wang et al. [96] applied a PEO(LiTFSI) interlayer between LAGP-PEO and lithium metal (Figure 8e), effectively suppressing interfacial side reactions and enabling initial discharge capacity of 161 mAh g−1 in assembled LiMn0.8Fe0.2PO4 all-solid-state batteries; Yu et al. [97] designed a “self-sacrificial” LiF interlayer on the surface of flexible LAGP/PPC electrolytes (Figure 8f), achieving 92.3% capacity retention after 100 cycles of LiFePO4/Li cells at 55 °C. These studies employ 3D structural design, surface chemical modification, and innovative interfacial layers to systematically address high interface impedance and poor lithium metal compatibility in NASICON-type electrolytes, providing multidimensional solutions for the development of highly stable solid-state batteries.

3.3. Graphene and Carbon-Based Fillers

Graphene and other carbon-based materials, such as carbon nanotubes and carbon nanofibers, have also been widely explored as fillers in composite solid-state electrolytes. Their high electronic conductivity, large specific surface area, and tunable surface chemistry enable strong interactions with polymer matrices and inorganic phases. When appropriately functionalized, these carbon-based systems can facilitate ion dissociation, provide continuous ion transport pathways, and enhance the mechanical strength of the electrolyte. Moreover, their flexible structures improve interfacial contact with electrodes, thereby reducing interfacial resistance and contributing to more stable cycling performance. These unique characteristics make graphene and other carbon-derived materials attractive candidates for next-generation composite solid-state electrolytes.
When designing organic–inorganic hybrid solid-state electrolytes (OICSEs), the selection of active fillers requires a comprehensive balance of multiple properties—including ionic conductivity, chemical stability, mechanical strength, and electrode compatibility—to optimize overall system performance (Table 2). Key considerations in material selection include (1) the synergistic effect between ionic conductivity and polymer amorphization; (2) chemical stability (e.g., sulfides’ moisture sensitivity necessitating hydrophobic modification); (3) mechanical strength (e.g., three-dimensional skeleton designs achieving modulus > 1 GPa to suppress dendrite growth); and (4) interfacial compatibility with electrodes (e.g., in situ removal of Li2CO3 passivation layers on garnet fillers). Through multiscale optimization strategies, such as surface engineering, structural design, and interface regulation, it is possible to simultaneously enhance ionic conductivity, cycling stability, and safety, thereby providing a material foundation for the development of high-energy-density, long-life solid-state batteries.

4. Advanced Characterization Methods for Composite Solid Electrolytes

Amid the accelerating evolution of all-solid-state lithium batteries and modern energy storage solutions, the study of OICSEs urgently requires the use of advanced multi-scale and multidimensional characterization techniques to reveal their complex structures, inherent attributes, and interfacial dynamic processes. Consequently, a range of sophisticated characterization tools plays a crucial role and offers broad applicability in the investigation of OICSEs.

4.1. Time-of-Flight Secondary Ion Mass Spectrometry

TOF-SIMS, as a highly sensitive analytical technique, enables precise characterization of material surface properties and elemental composition. This method bombards the sample surface with an ion beam to generate secondary ions and then uses the time-of-flight differences of these ions to obtain chemical composition, molecular structure, and elemental distribution information. Thanks to its excellent spatial, temporal, and mass resolution, TOF-SIMS is widely applied in a detailed analysis of electrolyte/electrode interface chemical components, especially suitable for tracking the dynamic evolution of reaction products and intermediates on electrode/electrolyte surfaces. This multidimensional characterization capability provides critical support for revealing electrochemical reaction mechanisms.
Goodenough conducted investigations on the interface between lithium metal anodes and CPE-25LZP electrolytes using TOF-SIMS depth profiling and cross-sectional imaging [98]. After cycling tests of Li||Li symmetric cells, Zr ions were used as indicators of the bulk solid electrolyte, while the surface concentration distributions of CsLi2P and Li2ZrO4 species were mapped (Figure 9a). Three-dimensional sputtered volume characterization visually displayed the spatial distribution features of CsLi2P and Zr signals (Figure 9b). Notably, besides the surface enrichment of CsLi2P, both ions exhibited fragmented distributions, reflecting the granular nature of the solid electrolyte. Comparing the depth distribution profiles of CsLi2P and Li2ZrO4 in fresh composite films, films after lithium metal contact, and films post-cycling (Figure 9c) revealed a significant increase in these ion concentrations on the membrane surface after cycling or lithium interaction compared to the original film. To better clarify the chemical constituents present at the interface between lithium and solid electrolyte, the research team used high lateral-resolution cross-sectional imaging to detect the presence of CsLi2P and Li2ZrO4 species in the interfacial region, microscopically confirming the particulate characteristics of the solid electrolyte.
While TOF-SIMS provides outstanding spatial, temporal, and mass-resolving power in the study of electrochemical processes, its application still faces constraints and it still faces key challenges in visualizing complex dynamic multiphase electrochemical reactions. To enable extended secondary ion flight paths, the technique necessitates operation under high-vacuum conditions, which poses practical obstacles for intricate multiphase systems involving interfaces at the liquid–solid junction. Therefore, the creation of advanced measurement strategies and vacuum-adapted electrochemical microfluidic platforms is critical for a deeper investigation into electrochemical dynamics.

4.2. High-Angle Annular Dark Field Scanning Transmission Electron Microscopy

Chen et al. [51] employed HAADF-STEM and EELS techniques to confirm that lithium ions tend to enrich at the boundaries of polymer/polymer and polymer/inorganic phases. Through STEM-EELS lithium K-edge imaging analysis, lithium aggregation was observed around the periphery of polyacrylonitrile fibers. Further EELS spectral measurements were conducted on both the fiber interface region and the interior, validating this enrichment behavior (Figure 9d). The spectra indicate that the lithium-rich region surrounds the lithium lanthanum zirconium tantalum oxide (LLZTO) nanoparticles: the EELS spectrum in the LLZTO particle region (Region 1) shows broadened edge features in the 50 ~ 80 eV range (Figure 9e) without distinct characteristic signals; meanwhile, the La-N4,5 edge signal around 110 eV suggests that lithium ions inside LLZTO reside in a disordered chemical environment. In contrast, the interface region (Region 2) displays a lithium K-edge spectrum with a pronounced double-peak, which reflects a more homogeneous and structured lithium-ion coordination.
HR-TEM combined with HAADF-STEM can directly reveal the microscopic structure and compositional distribution within materials. This method allows for observing the spatial distribution, morphology, and interface states of different material components, thereby enhancing the understanding of the composite electrolyte’s internal structure. When coupled with Electron Energy Loss Spectroscopy (EELS), it allows for not only an analysis of elemental composition but also an investigation of electronic structure, laying the groundwork for the optimization and advancement of composite electrolyte characteristics. Specifically, HAADF-STEM achieves sub-nanometer high-resolution imaging, suitable for examining nanostructures and interface characteristics in composite electrolytes. Using electron tomography, this technique can also reconstruct three-dimensional internal structures and morphological features. Combined with EELS, enabling investigations into elemental dispersion and chemical valence, the method supplies in-depth information on electronic structures and bonding compositions. However, exposure to a high-energy electron beam can lead to damage in sensitive materials, inducing structural modifications or degradation that could influence the accuracy of observation results. Additionally, owing to the restricted imaging region, the results obtained may fail to represent the macroscopic characteristics and uniformity throughout the whole composite electrolyte material. Therefore, it is necessary to combine other characterization techniques to more comprehensively reveal the performance characteristics of composite electrolyte materials.

4.3. X-Ray Computed Tomography (CT)

CT as a nondestructive three-dimensional imaging technique enables in-depth analysis of the internal microstructure of OICSEs. Such insight is key to correlating microstructure with overall performance, optimizing material configurations, and enhancing the capabilities of solid-state batteries [99]. Using this technique, the 3D distribution of both organic and inorganic materials can be mapped, porosity measured, and pore and channel features determined. For example, Cui et al. [100] presented scanning electron microscopy (SEM) images of a self-supporting, highly porous p-LATP electrolyte about 260 μm thick and further used CT to reveal its porous microstructure (Figure 9f). The 3D reconstruction of p-LATP and corresponding 2D slice images in the x-y and x-z planes showed a uniform pore distribution, confirming that the prepared p-LATP possesses interconnected porous architecture. Additionally, 3D imaging helps study the effect of pore connectivity on ion conduction and observe defect evolution in composite electrolytes during use, providing insights into material lifespan and stability. CT technology thus plays a key supporting role in developing high-power solid-state batteries. Nevertheless, compared to nano-CT, conventional CT has lower resolution, making it more suitable for imaging larger-scale structures.
Currently, multiple standard techniques are applied to examine the physicochemical characteristics of OICSEs. These techniques include XPS to examine surface chemical states and elemental distribution; XRD and NDP for analyzing crystal structures and phase characteristics; thermal analysis methods such as DSC and TGA for examining the thermal properties and stability of electrolytes; and Raman spectroscopy and infrared spectroscopy for analyzing molecular vibration modes of the materials. The combined application of these diverse characterization methods enables researchers to comprehensively and meticulously elucidate the structure–performance relationships of OICSEs.

5. Summary and Perspective

Due to their inherent safety characteristics and high energy density, ASSLBs are considered leading options for future energy storage technologies. This review highlights recent advances in OICSEs, a pivotal area in solid electrolyte (SE) development, which benefits from combining the strengths of various electrolyte constituents. Discussion centers on the role of ceramic fillers in ionic conductivity, focusing on variables such as particle size, content, morphology, and dimensionality. Additionally, lithium-ion transport mechanisms and the roles of inert fillers and active fillers in enhancing OICSE performance are explored. Ultimately, the crucial role of advanced characterization methods in thoroughly examining the chemical makeup, microstructural features, and interfacial characteristics of OICSEs is emphasized. Despite considerable advances in OICSE research, ASSLBs equipped with such electrolytes still lack sufficient technological maturity to support practical applications and commercialization. In view of the challenges faced by ASSLBs using these electrolytes, this review summarizes and promotes possible future paths for research and development.
Despite previous advancements in the ionic conductivity of OICSEs, several key challenges remain. Currently, most OICSEs maintain ionic conductivities in the range of 10−4 to 10−2 mS cm−1. Future strategies to enhance ionic conductivity can focus on the following key directions: (1) combining fillers and electrolytes at different scales to design hierarchical porous structures to optimize ion transport channels from microscopic to macroscopic levels, thereby improving overall conductivity; (2) adjusting film thickness to decrease the path length for ion migration (common techniques include electrospinning, hot-press infiltration, and spinning-spray combined methods), ensuring both mechanical robustness and flexibility to mitigate lithium dendrite development; (3) applying materials bearing high concentrations of surface functional groups, or improving filler dispersion uniformity through surface modifications; (4) investigating novel materials exhibiting high lithium-ion transference numbers, such as polymers with enhanced ion mobility, conductive polymers, and innovative inorganic electrolytes; and (5) adopting intelligent engineering and computational techniques, utilizing computational modeling, machine learning, and other tools to assist molecular structure design of OICSEs. Due to the complexity of lithium-ion transport mechanisms, the precise migration patterns within OICSEs remain unclear, while understanding interfacial interactions is critical for enhancing ionic conductivity. In situ cutting-edge characterization approaches enable the observation of ion mobility and concentration shifts in OICSEs under diverse conditions. Together with molecular dynamics simulations and density functional theory computations, these approaches offer multiscale, in-depth insights into ion transport mechanisms, forming a foundation for predicting ion diffusion channels and revealing interfacial effects. Future research in this area could focus on further enhancing the fundamental understanding of the underlying mechanisms, particularly at the molecular and interfacial levels. Continued efforts are needed to develop advanced characterization techniques and computational modeling to better correlate structure with performance. In addition, the design of novel materials and hybrid systems may open new pathways to overcome current limitations in conductivity, stability, and scalability. Finally, translating laboratory advances into practical applications will require interdisciplinary collaboration, bridging materials science, engineering, and real-world implementation.

Author Contributions

X.X. wrote the main manuscript text. D.L. and S.H. carried out the task of collecting literature. F.W., Y.M., and Q.X. all revised the draft. All authors reviewed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (22479094).

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no competing interests.

References

  1. Li, M.; Lu, J.; Chen, Z.; Amine, K. 30 years of lithium-ion batteries. Adv. Mater. 2018, 30, 1800561. [Google Scholar] [CrossRef] [PubMed]
  2. Tarascon, J.-M.; Armand, M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, F.-Q.; Wang, W.-P.; Yin, Y.-X.; Zhang, S.-F.; Shi, J.-L.; Wang, L.; Zhang, X.-D.; Zheng, Y.; Zhou, J.-J.; Li, L. Upgrading traditional liquid electrolyte via in situ gelation for future lithium metal batteries. Sci. Adv. 2018, 4, 5383. [Google Scholar] [CrossRef]
  4. Wang, S.-H.; Yue, J.; Dong, W.; Zuo, T.-T.; Li, J.-Y.; Liu, X.; Zhang, X.-D.; Liu, L.; Shi, J.-L.; Yin, Y.-X. Tuning wettability of molten lithium via a chemical strategy for lithium metal anodes. Nat. Commun. 2019, 10, 4930. [Google Scholar] [CrossRef] [PubMed]
  5. Harper, G.; Sommerville, R.; Kendrick, E.; Driscoll, L.; Slater, P.; Stolkin, R.; Walton, A.; Christensen, P.; Heidrich, O.; Lambert, S. Recycling lithium-ion batteries from electric vehicles. Nature 2019, 575, 75–86. [Google Scholar] [CrossRef]
  6. Wu, J.; Zheng, M.; Liu, T.; Wang, Y.; Liu, Y.; Nai, J.; Zhang, L.; Zhang, S.; Tao, X. Direct recovery: A sustainable recycling technology for spent lithium-ion battery. Energy Storage Mater. 2024, 54, 120–134. [Google Scholar] [CrossRef]
  7. Li, M.; An, H.; Song, Y.; Liu, Q.; Wang, J.; Huo, H.; Lou, S.; Wang, J. Ion-dipole-interaction-induced encapsulation of free residual solvent for long-cycle solid-state lithium metal batteries. J. Am. Chem. Soc. 2024, 145, 25632–25642. [Google Scholar] [CrossRef]
  8. Gicha, B.B.; Tufa, L.T.; Nwaji, N.; Hu, X.; Lee, J. Advances in all-solid-state lithium-sulfur batteries for commercialization. Nano-Micro Lett. 2025, 16, 172. [Google Scholar] [CrossRef]
  9. Wang, L.; Liu, T.; Dai, A.; De Andrade, V.; Ren, Y.; Xu, W.; Lee, S.; Zhang, Q.; Gu, L.; Wang, S. Reaction inhomogeneity coupling with metal rearrangement triggers electrochemical degradation in lithium-rich layered cathode. Nat. Commun. 2021, 12, 5370. [Google Scholar] [CrossRef]
  10. Kasemchainan, J.; Zekoll, S.; Spencer Jolly, D.; Ning, Z.; Hartley, G.O.; Marrow, J.; Bruce, P.G. Critical stripping current leads to dendrite formation on plating in lithium anode solid electrolyte cells. Nat. Mater. 2019, 18, 1105–1111. [Google Scholar] [CrossRef]
  11. Wang, H.; Gao, H.; Chen, X.; Zhu, J.; Li, W.; Gong, Z.; Li, Y.; Wang, M.S.; Yang, Y. Linking the defects to the formation and growth of Li dendrite in all-solid-state batteries. Adv. Energy Mater. 2021, 11, 2102148. [Google Scholar] [CrossRef]
  12. Chen, Y.; Wang, Z.; Li, X.; Yao, X.; Wang, C.; Li, Y.; Xue, W.; Yu, D.; Kim, S.Y.; Yang, F. Li metal deposition and stripping in a solid-state battery via Coble creep. Nature 2020, 578, 251–255. [Google Scholar] [CrossRef]
  13. Wang, L.; Dai, A.; Xu, W.; Lee, S.; Cha, W.; Harder, R.; Liu, T.; Ren, Y.; Yin, G.; Zuo, P. Structural distortion induced by manganese activation in a lithium-rich layered cathode. J. Am. Chem. Soc. 2020, 142, 14966–14973. [Google Scholar] [CrossRef] [PubMed]
  14. Murugan, R.; Thangadurai, V.; Weppner, W. Fast lithium ion conduction in garnet-type Li7La3Zr2O12. Angew. Chem.-Int. Ed. 2007, 46, 7778. [Google Scholar] [CrossRef]
  15. Zheng, H.; Wu, S.; Tian, R.; Xu, Z.; Zhu, H.; Duan, H.; Liu, H. Intrinsic lithiophilicity of Li-garnet electrolytes enabling high-rate lithium cycling. Adv. Funct. Mater. 2020, 30, 1906189. [Google Scholar] [CrossRef]
  16. Wang, X.; Huang, S.; Guo, K.; Min, Y.; Xu, Q. Directed and continuous interfacial channels for optimized ion transport in solid-state electrolytes. Adv. Funct. Mater. 2023, 32, 2206976. [Google Scholar] [CrossRef]
  17. Huo, H.; Chen, Y.; Zhao, N.; Lin, X.; Luo, J.; Yang, X.; Liu, Y.; Guo, X.; Sun, X. In-situ formed Li2CO3-free garnet/Li interface by rapid acid treatment for dendrite-free solid-state batteries. Nano Energy 2019, 61, 119–125. [Google Scholar] [CrossRef]
  18. Zhou, D.; Tkacheva, A.; Tang, X.; Sun, B.; Shanmukaraj, D.; Li, P.; Zhang, F.; Armand, M.; Wang, G. Stable conversion chemistry-based lithium metal batteries enabled by hierarchical multifunctional polymer electrolytes with near-single ion conduction. Angew. Chem.-Int. Ed. 2019, 58, 6001–6006. [Google Scholar] [CrossRef]
  19. Xi, J.; Qiu, X.; Zheng, S.; Tang, X. Nanocomposite polymer electrolyte comprising PEO/LiClO4 and solid super acid: Effect of sulphated-zirconia on the crystallization kinetics of PEO. Polymer 2005, 46, 5702–5706. [Google Scholar] [CrossRef]
  20. Nakayama, M.; Wada, S.; Kuroki, S.; Nogami, M. Factors affecting cyclic durability of all-solid-state lithium polymer batteries using poly (ethylene oxide)-based solid polymer electrolytes. Energy Environ. Sci. 2010, 3, 1995–2002. [Google Scholar] [CrossRef]
  21. Lei, D.; He, Y.-B.; Huang, H.; Yuan, Y.; Zhong, G.; Zhao, Q.; Hao, X.; Zhang, D.; Lai, C.; Zhang, S. Cross-linked beta alumina nanowires with compact gel polymer electrolyte coating for ultra-stable sodium metal battery. Nat. Commun. 2019, 10, 4244. [Google Scholar] [CrossRef]
  22. Kim, H.W.; Manikandan, P.; Lim, Y.J.; Kim, J.H.; Nam, S.-c.; Kim, Y. Hybrid solid electrolyte with the combination of Li7La3Zr2O12 ceramic and ionic liquid for high voltage pseudo-solid-state Li-ion batteries. J. Mater. Chem. A 2016, 4, 17025–17032. [Google Scholar] [CrossRef]
  23. Pervez, S.A.; Kim, G.; Vinayan, B.P.; Cambaz, M.A.; Kuenzel, M.; Hekmatfar, M.; Fichtner, M.; Passerini, S. Overcoming the interfacial limitations imposed by the solid-solid interface in solid-state batteries using ionic liquid-based interlayers. Small 2020, 16, 2000279. [Google Scholar] [CrossRef] [PubMed]
  24. Fan, W.; Li, N.W.; Zhang, X.; Zhao, S.; Cao, R.; Yin, Y.; Xing, Y.; Wang, J.; Guo, Y.G.; Li, C. A dual-salt gel polymer electrolyte with 3D cross-linked polymer network for dendrite-free lithium metal batteries. Adv. Sci. 2018, 5, 1800559. [Google Scholar] [CrossRef]
  25. Gao, Y.; Yan, Z.; Gray, J.L.; He, X.; Wang, D.; Chen, T.; Huang, Q.; Li, Y.C.; Wang, H.; Kim, S.H. Polymer-inorganic solid-electrolyte interphase for stable lithium metal batteries under lean electrolyte conditions. Nat. Mater. 2019, 18, 384–389. [Google Scholar] [CrossRef] [PubMed]
  26. Wu, Q.; Fang, M.; Jiao, S.; Li, S.; Zhang, S.; Shen, Z.; Mao, S.; Mao, J.; Zhang, J.; Tan, Y. Phase regulation enabling dense polymer-based composite electrolytes for solid-state lithium metal batteries. Nat. Commun. 2024, 14, 6296. [Google Scholar] [CrossRef]
  27. Capiglia, C.; Mustarelli, P.; Quartarone, E.; Tomasi, C.; Magistris, A. Effects of nanoscale SiO2 on the thermal and transport properties of solvent-free, poly(ethylene oxide) (PEO)-based polymer electrolytes. Solid State Ionics 1999, 118, 73–79. [Google Scholar] [CrossRef]
  28. Hou, Y.; Sheng, Z.; Fu, C.; Kong, J.; Zhang, X. Hygroscopic holey graphene aerogel fibers enable highly efficient moisture capture, heat allocation and microwave absorption. Nat. Commun. 2023, 13, 1227. [Google Scholar] [CrossRef] [PubMed]
  29. Tang, W.; Tang, S.; Zhang, C.; Ma, Q.; Xiang, Q.; Yang, Y.W.; Luo, J. Simultaneously enhancing the thermal stability, mechanical modulus, and electrochemical performance of solid polymer electrolytes by incorporating 2D sheets. Adv. Energy Mater. 2018, 8, 1800866. [Google Scholar] [CrossRef]
  30. Zhang, Q.; Cao, D.; Ma, Y.; Natan, A.; Aurora, P.; Zhu, H. Sulfide-based solid-state electrolytes: Synthesis, stability, and potential for all-solid-state batteries. Adv. Mater. 2019, 31, 1901131. [Google Scholar] [CrossRef]
  31. Mo, Y.; Ong, S.P.; Ceder, G. First principles study of the Li10GeP2S12 lithium super ionic conductor material. Chem. Mater. 2012, 24, 15–17. [Google Scholar] [CrossRef]
  32. Jiang, Z.; Wang, S.; Chen, X.; Yang, W.; Yao, X.; Hu, X.; Han, Q.; Wang, H. Tape-casting Li0.34La0.56TiO3 ceramic electrolyte films permit high energy density of lithium-metal batteries. Adv. Mater. 2020, 32, 1906221. [Google Scholar] [CrossRef]
  33. Richards, W.D.; Miara, L.J.; Wang, Y.; Kim, J.C.; Ceder, G. Interface stability in solid-state batteries. Chem. Mater. 2016, 28, 266–273. [Google Scholar] [CrossRef]
  34. Shan, J.; Gu, R.; Xu, J.; Gong, S.; Guo, S.; Xu, Q.; Shi, P.; Min, Y. Heterojunction Ferroelectric Materials Enhance Ion Transport and Fast Charging of Polymer Solid Electrolytes for Lithium Metal Batteries. Adv. Energy Mater. 2025, 15, 2405220. [Google Scholar] [CrossRef]
  35. Zheng, Y.; Yao, Y.; Ou, J.; Li, M.; Luo, D.; Dou, H.; Li, Z.; Amine, K.; Yu, A.; Chen, Z. A review of composite solid-state electrolytes for lithium batteries: Fundamentals, key materials and advanced structures. Chem. Soc. Rev. 2020, 49, 8790–8839. [Google Scholar] [CrossRef]
  36. Vijayakumar, V.; Ghosh, M.; Asokan, K.; Sukumaran, S.B.; Kurungot, S.; Mindemark, J.; Brandell, D.; Winter, M.; Nair, J.R. 2D layered nanomaterials as fillers in polymer composite electrolytes for lithium batteries. Adv. Energy Mater. 2023, 13, 2203326. [Google Scholar] [CrossRef]
  37. Sun, J.; Liu, C.; Liu, H.; Li, J.; Zheng, P.; Zheng, Y.; Liu, Z. Advances in ordered architecture design of composite solid electrolytes for solid-state lithium batteries. Chem. Rec. 2023, 23, e202300044. [Google Scholar] [CrossRef]
  38. Zheng, F.; Li, C.; Li, Z.; Cao, X.; Luo, H.; Liang, J.; Zhao, X.; Kong, J. Advanced composite solid electrolytes for lithium batteries: Filler dimensional design and ion path optimization. Small 2023, 19, 2206355. [Google Scholar] [CrossRef]
  39. Chen, W.-P.; Duan, H.; Shi, J.-L.; Qian, Y.; Wan, J.; Zhang, X.-D.; Sheng, H.; Guan, B.; Wen, R.; Yin, Y.-X. Bridging interparticle Li+ conduction in a soft ceramic oxide electrolyte. J. Am. Chem. Soc. 2021, 143, 5717–5726. [Google Scholar] [CrossRef]
  40. Du, L.; Zhang, B.; Wang, X.; Dong, C.; Mai, L.; Xu, L. 3D frameworks in composite polymer electrolytes: Synthesis, mechanisms, and applications. Chem. Eng. J. 2023, 451, 138787. [Google Scholar] [CrossRef]
  41. Ranque, P.; Zagórski, J.; Devaraj, S.; Aguesse, F.; del Amo, J.M.L. Characterization of the interfacial Li-ion exchange process in a ceramic-polymer composite by solid state NMR. J. Mater. Chem. A 2021, 9, 17812–17820. [Google Scholar] [CrossRef]
  42. Chen, R.; Qu, W.; Guo, X.; Li, L.; Wu, F. The pursuit of solid-state electrolytes for lithium batteries: From comprehensive insight to emerging horizons. Mater. Horiz. 2016, 3, 487–516. [Google Scholar] [CrossRef]
  43. Zheng, J.; Tang, M.; Hu, Y.Y. Lithium ion pathway within Li7La3Zr2O12-polyethylene oxide composite electrolytes. Angew. Chem.-Int. Ed. 2016, 55, 12538–12542. [Google Scholar] [CrossRef] [PubMed]
  44. Zheng, J.; Hu, Y.-Y. New insights into the compositional dependence of Li-ion transport in polymer-ceramic composite electrolytes. ACS Appl. Mater. Interfaces 2018, 10, 4113–4120. [Google Scholar] [CrossRef]
  45. Zhu, P.; Yan, C.; Dirican, M.; Zhu, J.; Zang, J.; Selvan, R.K.; Chung, C.-C.; Jia, H.; Li, Y.; Kiyak, Y. Li0.33La0.557TiO3 ceramic nanofiber-enhanced polyethylene oxide-based composite polymer electrolytes for all-solid-state lithium batteries. J. Mater. Chem. A 2018, 6, 4279–4285. [Google Scholar] [CrossRef]
  46. Liu, X.; Peng, S.; Gao, S.; Cao, Y.; You, Q.; Zhou, L.; Jin, Y.; Liu, Z.; Liu, J. Electric-field-directed parallel alignment architecting 3D lithium-ion pathways within solid composite electrolyte. ACS Appl. Mater. Interfaces 2018, 10, 15691–15696. [Google Scholar] [CrossRef]
  47. Zhang, J.; Zhao, N.; Zhang, M.; Li, Y.; Chu, P.K.; Guo, X.; Di, Z.; Wang, X.; Li, H. Flexible and ion-conducting membrane electrolytes for solid-state lithium batteries: Dispersion of garnet nanoparticles in insulating polyethylene oxide. Nano Energy 2016, 28, 447–454. [Google Scholar] [CrossRef]
  48. Lin, D.; Liu, W.; Liu, Y.; Lee, H.R.; Hsu, P.-C.; Liu, K.; Cui, Y. High ionic conductivity of composite solid polymer electrolyte via in situ synthesis of monodispersed SiO2 nanospheres in poly (ethylene oxide). Nano Lett. 2016, 16, 459–465. [Google Scholar] [CrossRef]
  49. Wan, Z.; Lei, D.; Yang, W.; Liu, C.; Shi, K.; Hao, X.; Shen, L.; Lv, W.; Li, B.; Yang, Q.H. Low resistance–integrated all-solid-state battery achieved by Li7La3Zr2O12 nanowire upgrading polyethylene oxide (PEO) composite electrolyte and PEO cathode binder. Adv. Funct. Mater. 2019, 29, 1805301. [Google Scholar] [CrossRef]
  50. Xu, H.; Chien, P.-H.; Shi, J.; Li, Y.; Wu, N.; Liu, Y.; Hu, Y.-Y.; Goodenough, J.B. High-performance all-solid-state batteries enabled by salt bonding to perovskite in poly (ethylene oxide). Proc. Natl. Acad. Sci. USA 2019, 116, 18815–18821. [Google Scholar] [CrossRef] [PubMed]
  51. Hu, C.; Shen, Y.; Shen, M.; Liu, X.; Chen, H.; Liu, C.; Kang, T.; Jin, F.; Li, L.; Li, J. Superionic conductors via bulk interfacial conduction. J. Am. Chem. Soc. 2020, 142, 18035–18041. [Google Scholar] [CrossRef] [PubMed]
  52. Li, Z.; Huang, H.-M.; Zhu, J.-K.; Wu, J.-F.; Yang, H.; Wei, L.; Guo, X. Ionic conduction in composite polymer electrolytes: Case of PEO: Ga-LLZO composites. ACS Appl. Mater. Interfaces 2018, 11, 784–791. [Google Scholar] [CrossRef] [PubMed]
  53. Fenton, D. Complex of alkali metal ions with poly (ethylene oxide). Polymer 1973, 14, 589. [Google Scholar] [CrossRef]
  54. Lee, M.J.; Han, J.; Lee, K.; Lee, Y.J.; Kim, B.G.; Jung, K.-N.; Kim, B.J.; Lee, S.W. Elastomeric electrolytes for high-energy solid-state lithium batteries. Nature 2022, 601, 217–222. [Google Scholar] [CrossRef]
  55. Chen, G.; Zhang, F.; Zhou, Z.; Li, J.; Tang, Y. A flexible dual-ion battery based on PVDF-HFP-modified gel polymer electrolyte with excellent cycling performance and superior rate capability. Adv. Energy Mater. 2018, 8, 1801219. [Google Scholar] [CrossRef]
  56. Bag, S.; Zhou, C.; Kim, P.J.; Pol, V.G.; Thangadurai, V. LiF modified stable flexible PVDF-garnet hybrid electrolyte for high performance all-solid-state Li-S batteries. Energy Storage Mater. 2020, 24, 198–207. [Google Scholar] [CrossRef]
  57. Nicotera, I.; Coppola, L.; Oliviero, C.; Ranieri, G.A. Rheological properties and impedance spectroscopy of PMMA-PVDF blend and PMMA gel polymer electrolytes for advanced lithium batteries. Ionics 2005, 11, 87–94. [Google Scholar] [CrossRef]
  58. Subban, R.; Arof, A. Plasticiser interactions with polymer and salt in PVC-LiCF3SO3-DMF electrolytes. Eur. Polym. J 2004, 40, 1841–1847. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Lu, W.; Cong, L.; Liu, J.; Sun, L.; Mauger, A.; Julien, C.M.; Xie, H.; Liu, J. Cross-linking network based on Poly (ethylene oxide): Solid polymer electrolyte for room temperature lithium battery. J. Power Sources 2019, 420, 63–72. [Google Scholar] [CrossRef]
  60. Bouchet, R.; Maria, S.; Meziane, R.; Aboulaich, A.; Lienafa, L.; Bonnet, J.-P.; Phan, T.N.; Bertin, D.; Gigmes, D.; Devaux, D. Single-ion BAB triblock copolymers as highly efficient electrolytes for lithium-metal batteries. Nat. Mater. 2013, 12, 452–457. [Google Scholar] [CrossRef] [PubMed]
  61. MacGlashan, G.S.; Andreev, Y.G.; Bruce, P.G. Structure of the polymer electrolyte poly(ethylene oxide)6: LiAsF6. Nature 1999, 398, 792–794. [Google Scholar] [CrossRef]
  62. Itoh, T.; Horii, S.; Uno, T.; Kubo, M.; Yamamoto, O. Influence of hyperbranched polymer structure on ionic conductivity in composite polymer electrolytes of PEO/hyperbranched polymer/BaTiO3/Li salt system. Electrochim. Acta 2004, 50, 271–274. [Google Scholar] [CrossRef]
  63. D’Epifanio, A.; Serraino Fiory, F.; Licoccia, S.; Traversa, E.; Scrosati, B.; Croce, F. Metallic-lithium, LiFePO4-based polymer battery using PEO-ZrO2 nanocomposite polymer electrolyte. J. Appl. Electrochem. 2004, 34, 403–408. [Google Scholar] [CrossRef]
  64. Zhou, B.; Jo, Y.H.; Wang, R.; He, D.; Zhou, X.; Xie, X.; Xue, Z. Self-healing composite polymer electrolyte formed via supramolecular networks for high-performance lithium-ion batteries. J. Mater. Chem. A 2019, 7, 10354–10362. [Google Scholar] [CrossRef]
  65. Xu, Z.; Yang, T.; Chu, X.; Su, H.; Wang, Z.; Chen, N.; Gu, B.; Zhang, H.; Deng, W.; Zhang, H. Strong lewis acid-base and weak hydrogen bond synergistically enhancing ionic conductivity of poly(ethylene oxide)@ SiO2 electrolytes for a high rate capability Li-metal battery. ACS Appl. Mater. Interfaces 2020, 12, 10341–10349. [Google Scholar] [CrossRef] [PubMed]
  66. Didwal, P.N.; Singhbabu, Y.; Verma, R.; Sung, B.-J.; Lee, G.-H.; Lee, J.-S.; Chang, D.R.; Park, C.-J. An advanced solid polymer electrolyte composed of poly(propylene carbonate) and mesoporous silica nanoparticles for use in all-solid-state lithium-ion batteries. Energy Storage Mater. 2021, 37, 476–490. [Google Scholar] [CrossRef]
  67. Liu, W.; Lin, D.; Sun, J.; Zhou, G.; Cui, Y. Improved lithium ionic conductivity in composite polymer electrolytes with oxide-ion conducting nanowires. ACS Nano 2016, 10, 11407–11413. [Google Scholar] [CrossRef] [PubMed]
  68. Sheng, O.; Jin, C.; Luo, J.; Yuan, H.; Huang, H.; Gan, Y.; Zhang, J.; Xia, Y.; Liang, C.; Zhang, W. Mg2B2O5 nanowire enabled multifunctional solid-state electrolytes with high ionic conductivity, excellent mechanical properties, and flame-retardant performance. Nano Lett. 2018, 18, 3104–3112. [Google Scholar] [CrossRef]
  69. Hua, S.; Jing, M.X.; Han, C.; Yang, H.; Chen, H.; Chen, F.; Chen, L.L.; Ju, B.W.; Tu, F.Y.; Shen, X.Q. A novel titania nanorods-filled composite solid electrolyte with improved room temperature performance for solid-state Li-ion battery. Int. J. Energy Res. 2019, 43, 7296–7305. [Google Scholar] [CrossRef]
  70. Li, C.; Huang, Y.; Chen, C.; Feng, X.; Zhang, Z. High-performance polymer electrolyte membrane modified with isocyanate-grafted Ti3+ doped TiO2 nanowires for lithium batteries. Appl. Surf. Sci. 2021, 563, 150248. [Google Scholar] [CrossRef]
  71. Zhao, E.; Guo, Y.; Zhang, A.; Wang, H.; Xu, G. Polydopamine coated TiO2 nanofiber fillers for polyethylene oxide hybrid electrolytes for efficient and durable all solid state lithium ion batteries. Nanoscale 2022, 14, 890–897. [Google Scholar] [CrossRef] [PubMed]
  72. Wen, J.; Zhao, Q.; Jiang, X.; Ji, G.; Wang, R.; Lu, G.; Long, J.; Hu, N.; Xu, C. Graphene oxide enabled flexible PEO-based solid polymer electrolyte for all-solid-state lithium metal battery. ACS Appl. Energy Mater. 2021, 4, 3660–3669. [Google Scholar] [CrossRef]
  73. Chen, L.; Li, W.; Fan, L.Z.; Nan, C.W.; Zhang, Q. Intercalated electrolyte with high transference number for dendrite-free solid-state lithium batteries. Adv. Funct. Mater. 2019, 29, 1901047. [Google Scholar] [CrossRef]
  74. Shi, Y.; Li, B.; Zhu, Q.; Shen, K.; Tang, W.; Xiang, Q.; Chen, W.; Liu, C.; Luo, J.; Yang, S. MXene-based mesoporous nanosheets toward superior lithium ion conductors. Adv. Energy Mater. 2020, 10, 1903534. [Google Scholar] [CrossRef]
  75. Zhang, X.; Xie, J.; Shi, F.; Lin, D.; Liu, Y.; Liu, W.; Pei, A.; Gong, Y.; Wang, H.; Liu, K. Vertically aligned and continuous nanoscale ceramic-polymer interfaces in composite solid polymer electrolytes for enhanced ionic conductivity. Nano Lett. 2018, 18, 3829–3838. [Google Scholar] [CrossRef] [PubMed]
  76. Lin, D.; Yuen, P.Y.; Liu, Y.; Liu, W.; Liu, N.; Dauskardt, R.H.; Cui, Y. A silica-aerogel-reinforced composite polymer electrolyte with high ionic conductivity and high modulus. Adv. Mater. 2018, 30, 1802661. [Google Scholar] [CrossRef]
  77. Zhang, Z.; Huang, Y.; Gao, H.; Li, C.; Huang, J.; Liu, P. 3D glass fiber cloth reinforced polymer electrolyte for solid-state lithium metal batteries. J. Membr. Sci. 2021, 621, 118940. [Google Scholar] [CrossRef]
  78. Zhang, Z.; Wang, Q.; Li, Z.; Jiang, Y.; Zhao, B.; Han, X. Well-aligned BaTiO3 nanofibers via solution blow spinning and their application in lithium composite solid-state electrolyte. Mater. Express 2019, 9, 993–1000. [Google Scholar] [CrossRef]
  79. Zhao, Y.; Wu, C.; Peng, G.; Chen, X.; Yao, X.; Bai, Y.; Wu, F.; Chen, S.; Xu, X. A new solid polymer electrolyte incorporating Li10GeP2S12 into a polyethylene oxide matrix for all-solid-state lithium batteries. J. Power Sources 2016, 301, 47–53. [Google Scholar] [CrossRef]
  80. Chen, S.; Wang, J.; Zhang, Z.; Wu, L.; Yao, L.; Wei, Z.; Deng, Y.; Xie, D.; Yao, X.; Xu, X. In-situ preparation of poly(ethylene oxide)/Li3PS4 hybrid polymer electrolyte with good nanofiller distribution for rechargeable solid-state lithium batteries. J. Power Sources 2018, 387, 72–80. [Google Scholar] [CrossRef]
  81. Li, Y.; Arnold, W.; Thapa, A.; Jasinski, J.B.; Sumanasekera, G.; Sunkara, M.; Druffel, T.; Wang, H. Stable and flexible sulfide composite electrolyte for high-performance solid-state lithium batteries. ACS Appl. Mater. Interfaces 2020, 12, 42653–42659. [Google Scholar] [CrossRef]
  82. Liu, S.; Zhou, L.; Han, J.; Wen, K.; Guan, S.; Xue, C.; Zhang, Z.; Xu, B.; Lin, Y.; Shen, Y. Super long-cycling all-solid-state battery with thin Li6PS5Cl-based electrolyte. Adv. Energy Mater. 2022, 12, 2200660. [Google Scholar] [CrossRef]
  83. Luo, S.; Wang, Z.; Fan, A.; Liu, X.; Wang, H.; Ma, W.; Zhu, L.; Zhang, X. A high energy and power all-solid-state lithium battery enabled by modified sulfide electrolyte film. J. Power Sources 2021, 485, 229325. [Google Scholar] [CrossRef]
  84. Su, Y.; Zhang, X.; Du, C.; Luo, Y.; Chen, J.; Yan, J.; Zhu, D.; Geng, L.; Liu, S.; Zhao, J. An all-solid-state battery based on sulfide and PEO composite electrolyte. Small 2022, 18, 2202069. [Google Scholar] [CrossRef]
  85. Choi, J.-H.; Lee, C.-H.; Yu, J.-H.; Doh, C.-H.; Lee, S.-M. Enhancement of ionic conductivity of composite membranes for all-solid-state lithium rechargeable batteries incorporating tetragonal Li7La3Zr2O12 into a polyethylene oxide matrix. J. Power Sources 2015, 274, 458–463. [Google Scholar] [CrossRef]
  86. Huang, Z.; Pang, W.; Liang, P.; Jin, Z.; Grundish, N.; Li, Y.; Wang, C.-A. A dopamine modified Li6.4La3Zr1.4Ta0.6O12/PEO solid-state electrolyte: Enhanced thermal and electrochemical properties. J. Mater. Chem. A 2019, 7, 16425–16436. [Google Scholar] [CrossRef]
  87. Li, W.; Sun, C.; Jin, J.; Li, Y.; Chen, C.; Wen, Z. Realization of the Li+ domain diffusion effect via constructing molecular brushes on the LLZTO surface and its application in all-solid-state lithium batteries. J. Mater. Chem. A 2019, 7, 27304–27312. [Google Scholar] [CrossRef]
  88. Wu, L.; Wang, Y.; Tang, M.; Liang, Y.; Lin, Z.; Ding, P.; Zhang, Z.; Wang, B.; Liu, S.; Li, L. Lithium-ion transport enhancement with bridged ceramic-polymer interface. Energy Storage Mater. 2023, 58, 40–47. [Google Scholar] [CrossRef]
  89. Dai, J.; Fu, K.; Gong, Y.; Song, J.; Chen, C.; Yao, Y.; Pastel, G.; Zhang, L.; Wachsman, E.; Hu, L. Flexible solid-state electrolyte with aligned nanostructures derived from wood. ACS Mater. Lett. 2019, 1, 354–361. [Google Scholar] [CrossRef]
  90. OV, S.; Elsin Abraham, S.; Ramaswamy, M. Free-standing and flexible garnet-PVDF ceramic polymer electrolyte membranes for solid-state batteries. Energy Fuels 2023, 37, 2401–2409. [Google Scholar] [CrossRef]
  91. Yu, X.; Liu, Y.; Goodenough, J.B.; Manthiram, A. Rationally designed PEGDA-LLZTO composite electrolyte for solid-state lithium batteries. ACS Appl. Mater. Interfaces 2021, 13, 30703–30711. [Google Scholar] [CrossRef]
  92. Wang, Y.-J.; Pan, Y.; Kim, D. Conductivity studies on ceramic Li1.3Al0.3Ti1.7(PO4)3-filled PEO-based solid composite polymer electrolytes. J. Power Sources 2006, 159, 690–701. [Google Scholar] [CrossRef]
  93. Jin, Y.; Zong, X.; Zhang, X.; Jia, Z.; Xie, H.; Xiong, Y. Constructing 3D Li+-percolated transport network in composite polymer electrolytes for rechargeable quasi-solid-state lithium batteries. Energy Storage Mater. 2022, 49, 433–444. [Google Scholar] [CrossRef]
  94. Wang, G.; Liu, H.; Liang, Y.; Wang, C.; Fan, L.-Z. Composite polymer electrolyte with three-dimensional ion transport channels constructed by NaCl template for solid-state lithium metal batteries. Energy Storage Mater. 2022, 45, 1212–1219. [Google Scholar] [CrossRef]
  95. Jin, Y.; Zong, X.; Zhang, X.; Liu, C.; Li, D.; Jia, Z.; Li, G.; Zhou, X.; Wei, J.; Xiong, Y. Interface regulation enabling three-dimensional Li1.3Al0.3Ti1.7(PO4)3-reinforced composite solid electrolyte for high-performance lithium batteries. J. Power Sources 2021, 501, 230027. [Google Scholar] [CrossRef]
  96. Wang, C.; Yang, Y.; Liu, X.; Zhong, H.; Xu, H.; Xu, Z.; Shao, H.; Ding, F. Suppression of lithium dendrite formation by using LAGP-PEO (LiTFSI) composite solid electrolyte and lithium metal anode modified by PEO (LiTFSI) in all-solid-state lithium batteries. ACS Appl. Mater. Interfaces 2017, 9, 13694–13702. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, L.; Hu, S.; Su, J.; Huang, T.; Yu, A. Self-sacrificed interface-based on the flexible composite electrolyte for high-performance all-solid-state lithium batteries. ACS Appl. Mater. Interfaces 2019, 11, 42715–42721. [Google Scholar] [CrossRef]
  98. Wu, N.; Chien, P.-H.; Li, Y.; Dolocan, A.; Xu, H.; Xu, B.; Grundish, N.S.; Jin, H.; Hu, Y.-Y.; Goodenough, J.B. Fast Li+ conduction mechanism and interfacial chemistry of a NASICON/polymer composite electrolyte. J. Am. Chem. Soc. 2020, 142, 2497–2505. [Google Scholar] [CrossRef]
  99. Wang, Y.; Ju, J.; Dong, S.; Yan, Y.; Jiang, F.; Cui, L.; Wang, Q.; Han, X.; Cui, G. Facile design of sulfide-based all solid-state lithium metal battery: In situ polymerization within self-supported porous argyrodite skeleton. Adv. Funct. Mater. 2021, 31, 2101523. [Google Scholar] [CrossRef]
  100. Yan, Y.; Ju, J.; Dong, S.; Wang, Y.; Huang, L.; Cui, L.; Jiang, F.; Wang, Q.; Zhang, Y.; Cui, G. In situ polymerization permeated three-dimensional Li+-percolated porous oxide ceramic framework boosting all solid-state lithium metal battery. Adv. Sci. 2023, 8, 2003887. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Composite solid electrolytes with different filler dimensional designs. Reproduced with permission from Kong, J, Small; published by WILEY, 2024 [38].
Figure 1. Composite solid electrolytes with different filler dimensional designs. Reproduced with permission from Kong, J, Small; published by WILEY, 2024 [38].
Processes 13 02795 g001
Figure 2. A schematic diagram illustrating the conduction mechanisms in the two regions of polymer electrolytes. Reproduced with permission from Wu, F, Mater. Horiz.; published by Royal Society of Chemistry, 2016 [42].
Figure 2. A schematic diagram illustrating the conduction mechanisms in the two regions of polymer electrolytes. Reproduced with permission from Wu, F, Mater. Horiz.; published by Royal Society of Chemistry, 2016 [42].
Processes 13 02795 g002
Figure 3. (a) Illustration of the interaction mechanism among PEO chains and Al2O3 surface groups. Reproduced with permission from Croce, F, J. Appl. Electrochem; published by Springer, 2004 [63]. (b) Process flow for preparing SiO2-UPy and schematic of the SHCPE architecture based on supramolecular networks. Reproduced with permission from Xue, Z, J. Mater. Chem. A; published by Royal Society of Chemistry, 2019 [64]. (c) The brief mechanism of strong Lewis acid–base interaction between PEO chains, SiO2, nanoparticles, and LiClO4. Reproduced with permission from Zhang H, ACS Appl. Mater. Interfaces; published by American Chemical Society, 2020 [65]. (d) Conceptual schematic of composite solid polymer electrolytes. Reproduced with permission from Park, C.-J, Energy Storage Mater.; published by Elsevier, 2021 [66].
Figure 3. (a) Illustration of the interaction mechanism among PEO chains and Al2O3 surface groups. Reproduced with permission from Croce, F, J. Appl. Electrochem; published by Springer, 2004 [63]. (b) Process flow for preparing SiO2-UPy and schematic of the SHCPE architecture based on supramolecular networks. Reproduced with permission from Xue, Z, J. Mater. Chem. A; published by Royal Society of Chemistry, 2019 [64]. (c) The brief mechanism of strong Lewis acid–base interaction between PEO chains, SiO2, nanoparticles, and LiClO4. Reproduced with permission from Zhang H, ACS Appl. Mater. Interfaces; published by American Chemical Society, 2020 [65]. (d) Conceptual schematic of composite solid polymer electrolytes. Reproduced with permission from Park, C.-J, Energy Storage Mater.; published by Elsevier, 2021 [66].
Processes 13 02795 g003
Figure 4. (a) Illustration depicting Li-ion conduction assisted by nanoparticle and nanowire inclusions. Reproduced with permission from Zhang, W, Nano Lett.; published by American Chemical Society, 2018 [67]. (b) Schematic illustration of lithium-ion transport in Mg2B2O5 enhanced OICSEs. Reproduced with permission from Sheng, X. Q, Int. J. Energy Res.; published by Wiley, 2019 [68]. (c) Schematic illustration of TiO2 modified with TDI and OICSE preparation. Reproduced with permission from Xu, G, Nanoscale; published by Royal Society of Chemistry, 2022 [70]. (d) Schematic illustration of the OICSE preparation process. Reproduced with permission from Xu, C, ACS. Appl. Energy Mater.; published by American Chemical Society, 2021 [71].
Figure 4. (a) Illustration depicting Li-ion conduction assisted by nanoparticle and nanowire inclusions. Reproduced with permission from Zhang, W, Nano Lett.; published by American Chemical Society, 2018 [67]. (b) Schematic illustration of lithium-ion transport in Mg2B2O5 enhanced OICSEs. Reproduced with permission from Sheng, X. Q, Int. J. Energy Res.; published by Wiley, 2019 [68]. (c) Schematic illustration of TiO2 modified with TDI and OICSE preparation. Reproduced with permission from Xu, G, Nanoscale; published by Royal Society of Chemistry, 2022 [70]. (d) Schematic illustration of the OICSE preparation process. Reproduced with permission from Xu, C, ACS. Appl. Energy Mater.; published by American Chemical Society, 2021 [71].
Processes 13 02795 g004
Figure 5. (a) Preparation schematic of GO-PEO hybrid solid polymer electrolyte alongside assembly steps for solid-state LiFePO4||GO-PEO||Li battery. Reproduced with permission from Zhang, Q, Adv. Funct. Mater.; published by Wiley, 2019 [72]. (b) Digital photographs and matching top-view and cross-sectional SEM images of Ni foam and Li-Ni composite. Reproduced with permission from Yang, S, Adv. Energy Mater.; published by Wiley, 2020 [73]. (c) Schematics of composite solid polymer electrolyte featuring three geometric configurations of the ceramic–polymer interface. Reproduced with permission from Cui, Y, Adv. Mater.; published by Wiley, 2018 [75]. (d) Schematic representation of the fabrication procedure for SiO2-aerogel-enhanced CPE. Reproduced with permission from Liu, P, J. Membr. Sci.; published by Elsevier, 2021 [76].
Figure 5. (a) Preparation schematic of GO-PEO hybrid solid polymer electrolyte alongside assembly steps for solid-state LiFePO4||GO-PEO||Li battery. Reproduced with permission from Zhang, Q, Adv. Funct. Mater.; published by Wiley, 2019 [72]. (b) Digital photographs and matching top-view and cross-sectional SEM images of Ni foam and Li-Ni composite. Reproduced with permission from Yang, S, Adv. Energy Mater.; published by Wiley, 2020 [73]. (c) Schematics of composite solid polymer electrolyte featuring three geometric configurations of the ceramic–polymer interface. Reproduced with permission from Cui, Y, Adv. Mater.; published by Wiley, 2018 [75]. (d) Schematic representation of the fabrication procedure for SiO2-aerogel-enhanced CPE. Reproduced with permission from Liu, P, J. Membr. Sci.; published by Elsevier, 2021 [76].
Processes 13 02795 g005
Figure 7. (a) Schematic illustration of an integrated LiFPO4/CSE/Li battery [49] Reproduced with permission from Wang, C. A. J. Mater. Chem. A; published by Royal Society of Chemistry, 2019. (b) Diagram depicting the architecture of MB-LLZTO-CPE and the effect of lithium-ion domain diffusion. Reproduced with permission from Li, L. J. Mater. Chem. A; published by Royal Society of Chemistry, 2019 [87]. (c) Illustration of the synthesis process for an OICSE that mediates interaction between the polymer matrix and ceramic fillers. Reproduced with permission from Hu, L. ACS Mater. Lett.; published by American Chemical Society, 2019 [88]. (d) Visual representation of a mesoporous LLZO membrane with multiscale alignment, combined with PEO polymer. Reproduced with permission from Ramaswamy, M. ENERG·FUEL; published by American Chemical Society, 2023 [89].
Figure 7. (a) Schematic illustration of an integrated LiFPO4/CSE/Li battery [49] Reproduced with permission from Wang, C. A. J. Mater. Chem. A; published by Royal Society of Chemistry, 2019. (b) Diagram depicting the architecture of MB-LLZTO-CPE and the effect of lithium-ion domain diffusion. Reproduced with permission from Li, L. J. Mater. Chem. A; published by Royal Society of Chemistry, 2019 [87]. (c) Illustration of the synthesis process for an OICSE that mediates interaction between the polymer matrix and ceramic fillers. Reproduced with permission from Hu, L. ACS Mater. Lett.; published by American Chemical Society, 2019 [88]. (d) Visual representation of a mesoporous LLZO membrane with multiscale alignment, combined with PEO polymer. Reproduced with permission from Ramaswamy, M. ENERG·FUEL; published by American Chemical Society, 2023 [89].
Processes 13 02795 g007
Figure 8. (a) Schematic illustration for the preparation of 3D composite fiber network-reinforced CPE. Reproduced with permission from Fan, L. Z. Energy Storage Mater.; published by Elsevier, 2022 [93]. (b) Schematic diagram of the fabrication of CPE−3D and long-term cycling performance of Li||PEO||Li, Li||CPE-R||Li, and Li||CPE−3D||Li symmetric batteries at 0.2 mA cm−2. (c) Top and cross-sectional images of the 3D porous LATP framework fabricated with varying NaCl template mass fractions. Reproduced with permission from Xiong, Y. J. Power Sources; published by Elsevier, 2021 [94]. (d) Illustration of the fabrication process and working principle of the 3D LATP@PMMA-PVDF composite electrolyte used in NCM811/Li solid-state batteries. Reproduced with permission from Ding, F. ACS Appl. Mater.; published by American Chemical Society, 2017 [95]. (e) All-solid-state Li-PEO (LiTFSI)||LAGP-PEO (LiTFSI)||LiMFP cells. Reproduced with permission from Yu, A. ACS Appl. Mater. Interfaces; published by American Chemical Society, 2019 [96]. (f) Schematics of the all-solid-state battery assembly. Reproduced with permission from Goodenough, J. B. J. Am. Chem. Soc.; published by American Chemical Society, 2020 [97].
Figure 8. (a) Schematic illustration for the preparation of 3D composite fiber network-reinforced CPE. Reproduced with permission from Fan, L. Z. Energy Storage Mater.; published by Elsevier, 2022 [93]. (b) Schematic diagram of the fabrication of CPE−3D and long-term cycling performance of Li||PEO||Li, Li||CPE-R||Li, and Li||CPE−3D||Li symmetric batteries at 0.2 mA cm−2. (c) Top and cross-sectional images of the 3D porous LATP framework fabricated with varying NaCl template mass fractions. Reproduced with permission from Xiong, Y. J. Power Sources; published by Elsevier, 2021 [94]. (d) Illustration of the fabrication process and working principle of the 3D LATP@PMMA-PVDF composite electrolyte used in NCM811/Li solid-state batteries. Reproduced with permission from Ding, F. ACS Appl. Mater.; published by American Chemical Society, 2017 [95]. (e) All-solid-state Li-PEO (LiTFSI)||LAGP-PEO (LiTFSI)||LiMFP cells. Reproduced with permission from Yu, A. ACS Appl. Mater. Interfaces; published by American Chemical Society, 2019 [96]. (f) Schematics of the all-solid-state battery assembly. Reproduced with permission from Goodenough, J. B. J. Am. Chem. Soc.; published by American Chemical Society, 2020 [97].
Processes 13 02795 g008
Figure 9. (a) Normalized TOF-SIMS depth profiles of three species of interest. (b) Three-dimensional view of the sputtered volume in panel a. Only CsLi2P and Zr are shown for clarity. (c) Depth profiles derived from cycled, handled, and pristine samples. Reproduced with permission from Li, J. J. Am. Chem. Soc.; published by American Chemical Society, 2020 [51]. (d) HAADF-TEM images of PAN/LiClO4 and the PAN/LiClO4. (e) Selected region EELS spectra covering organic particles, organic-to-organic interfaces, polymer phase, and polymer-to-inorganic interfaces. Reproduced with permission from Cui, G. Adv. Funct. Mater.; published by Wiley, 2021 [99]. (f) Cross-sectional SEM view, 3D reconstruction, and 2D sectional images of p-LATP taken from x-y and x-z planes. Reproduced with permission from Cui, G. Adv. Sci.; published by Wiley, 2023 [100].
Figure 9. (a) Normalized TOF-SIMS depth profiles of three species of interest. (b) Three-dimensional view of the sputtered volume in panel a. Only CsLi2P and Zr are shown for clarity. (c) Depth profiles derived from cycled, handled, and pristine samples. Reproduced with permission from Li, J. J. Am. Chem. Soc.; published by American Chemical Society, 2020 [51]. (d) HAADF-TEM images of PAN/LiClO4 and the PAN/LiClO4. (e) Selected region EELS spectra covering organic particles, organic-to-organic interfaces, polymer phase, and polymer-to-inorganic interfaces. Reproduced with permission from Cui, G. Adv. Funct. Mater.; published by Wiley, 2021 [99]. (f) Cross-sectional SEM view, 3D reconstruction, and 2D sectional images of p-LATP taken from x-y and x-z planes. Reproduced with permission from Cui, G. Adv. Sci.; published by Wiley, 2023 [100].
Processes 13 02795 g009
Table 1. Comparison of the benefits and limitations of inert materials with different dimensionalities.
Table 1. Comparison of the benefits and limitations of inert materials with different dimensionalities.
Inert FillersExampleBenefitsLimitationsRef.
0D particlesSiO2/Al2O3/ZrO2Strong mechanical integrity and high hardness
Superior chemical resistance
Reliable heat stability
Poor ionic conductivity
(<10−3 mS cm−1)
Insufficient contact at the interface of electrolyte and electrodes
J. Mater. Chem. A (2019) [64]
1D nanofiberHNTs
TiO2/Y2O3 Nanowires
Relatively high specific surface area
Superior chemical resistance
Good mechanical properties
Poor ionic conductivity
Intricate fabrication method
HNTs with brittle structure
Appl. Surf. Sci. (2021) [70]
2D NanosheetMXene-Ti3C2, GO, BNLarge specific surface area
Multifunctional surface functional group
Good mechanical properties
Large differences in ionic conductivity
High preparation process and cost
Adv. Energy Mater. (2020) [74]
3D Network3D-Al2O3, 3D-SiO2High specific surface area
Excellent mechanical properties and thermal stability
Low ionic conductivity
(<10−1 mS cm−1)
Complex preparation process
Adv. Mater. (2018) [77]
Table 2. Electrochemical properties of OICSEs with different types of fillers.
Table 2. Electrochemical properties of OICSEs with different types of fillers.
Fillersσ (mS cm−1)TLi+EW (V)Ref.
0D-SiO20.14 (30 °C)0.24.55J. Mater. Chem. A (2019) [64]
1D-TiO20.1 (RT)0.54.95Appl. Surf. Sci. (2021) [71]
2D-Mxene-mSiO20.46 (25 °C)0.614.8Adv. Energy Mater. (2020) [74]
3D-Al2O30.58 (RT)0.454.5Adv. Mater. (2018) [77]
Sulfide-type materials Li10GeP2S120.42 (30 °C)0.874.6J. Power Sources (2016) [79]
Garnet-type materials Li7La3Zr2O120.31 (RT)0.705.4ACS Appl. Mater. Interfaces (2021) [90]
NASICON-type materials Li1.3Al0.3Ti1.7(PO4)31.06 (25 °C)0.824.86Energy Storage Mater. (2022) [94]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xu, X.; Lu, D.; Huang, S.; Wang, F.; Min, Y.; Xu, Q. Multiscale Insights into Inorganic Filler Regulation, Ion Transport Mechanisms, and Characterization Advances in Composite Solid-State Electrolytes. Processes 2025, 13, 2795. https://doi.org/10.3390/pr13092795

AMA Style

Xu X, Lu D, Huang S, Wang F, Min Y, Xu Q. Multiscale Insights into Inorganic Filler Regulation, Ion Transport Mechanisms, and Characterization Advances in Composite Solid-State Electrolytes. Processes. 2025; 13(9):2795. https://doi.org/10.3390/pr13092795

Chicago/Turabian Style

Xu, Xinhao, Dingyuan Lu, Sipeng Huang, Fuming Wang, Yulin Min, and Qunjie Xu. 2025. "Multiscale Insights into Inorganic Filler Regulation, Ion Transport Mechanisms, and Characterization Advances in Composite Solid-State Electrolytes" Processes 13, no. 9: 2795. https://doi.org/10.3390/pr13092795

APA Style

Xu, X., Lu, D., Huang, S., Wang, F., Min, Y., & Xu, Q. (2025). Multiscale Insights into Inorganic Filler Regulation, Ion Transport Mechanisms, and Characterization Advances in Composite Solid-State Electrolytes. Processes, 13(9), 2795. https://doi.org/10.3390/pr13092795

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

Article Metrics

Back to TopTop