J. Electrochem. Sci. Technol Search

CLOSE


J. Electrochem. Sci. Technol > Epub ahead of print
Youn, Kim, Kim, Kim, Kim, and Kim: SiOx-based High-capacity Anode Materials for Li-ion Batteries

Abstract

Si-based anode materials have drawn much attention because of their higher specific capacity compared to graphite currently used in the anode of commercial lithium-ion batteries (LIBs). However, their practical use is still limited because of poor cycle performance derived from a high volume expansion rate of ~300% during cycling. Silicon suboxides (SiOx) are under intensive investigation as a very promising anode material for next-generation LIBs, with the highest cycle performance among Si-based ones thanks to the formation of a buffer phase during the first cycle of lithiation. This mini-review covers research progresses in SiOx-based high-capacity anode materials. The structural models and electrochemical characteristics of SiOx are presented, as well as efforts to improve its electrochemical performance.

INTRODUCTION

Lithium-ion batteries (LIBs) are one of the most promising energy storage devices for use as power sources in electric vehicles (EV) and energy storage systems (ESS) [18]. Commercial LIBs generally adopt graphite as the anode material because of its stable cycle retention and high initial Coulombic efficiency (ICE) [9,10]. However, the low theoretical capacity of graphite (372 mAh g–1) poses a problem for the EV and ESS industries, which demand high energy densities [1113]. Silicon (Si) has attracted extensive attention as an alternative anode material, because its theoretical capacity (3580 mAh g–1) is almost ten times that of graphite [14,15]. Nevertheless, the cycle life of Si electrodes is unsatisfactory for practical use owing to their high volume expansion rate (~300%) during the lithiation process, which causes pulverization, loss of electrical contact, and delamination of the electrode material [1619]. These drawbacks may be overcome using Si/carbon composites, silicon suboxides (SiOx), and Si/alloy composites [2022].
Compared to other proposed Si-based materials, SiOx shows much improved cycle performance, owing to the buffering role of lithium oxide and Li–Si–O ternary phases produced during the initial lithiation process, while it still shows 3–4 times greater capacity than that of graphite. With these advantages, SiOx anode material has been applied in commercial LIBs as the first Si-based anode material, with contents including up to 10 wt% [2330]. However, its low ICE, inferior long-term cycle performance compared to graphite, and low electrical conductivity prevent the adoption of pure SiOx without graphite in commercial LIBs [31,32]. The low ICE arises from the initial consumption of lithium ion (Li+) to form lithium oxide and Li–Si–O ternary phases, which causes a severe drop of energy density in the full cell because there is a limited amount of Li+. Researchers attempted to passivate the SiOx material or electrode (i.e., by reacting SiOx with lithium before cell production) to elevate the ICE. Nevertheless, the long-term cycle performance of SiOx electrodes was still poorer than that of graphite, mainly because of its high volume expansion rate (160% vs. 13% for graphite) [28,33]. Nanostructured SiOx materials such as nano-rod SiOx are promising for improving the long-term cycling performance by controlling the direction of volume expansion [34]. The low electrical conductivity of SiOx originates from the regions of SiO2, which is considered an insulator [26]. This low electrical conductivity can degrade the electrochemical performance and result in a poor rate capability. The combination of SiOx with carbon is an effective method for enhancing the electrical conductivity. For example, carbon coating on a SiO material increased the electrical conductivity and maintained the electrical contact network of each particle [32]. However, despite these efforts, it remains difficult to fully replace graphite with SiOx as the anode material for commercial LIBs due to problems related to electrochemical performance and production cost.
This mini-review introduces structural models proposed for SiOx and its electrochemical characteristics. Research progress in improving the electrochemical performance of SiOx is also discussed, along with current issues in this area.

STRUCTURAL MODELS OF SiOx

An accurate assessment of the structure of SiOx is complicated by its non-stoichiometric nature [31,35]. For the stoichiometric amorphous silicon monoxide (a-SiO), Brady was the first to propose a structure model using a radial distribution function (RDF) derived from the X-ray diffraction (XRD) pattern [36]. Because the RDF data indicated that each Si atom is surrounded by 3.7 nearest Si atoms, Brady suggested a random mixture model (RM) by assuming that SiO is a simple mixture of Si and SiO2 in an atomic ratio of 1:1 [37]. In contrast, the random bonding model (RB) assumes that Si-Si and Si-O bonds are randomly networked in the same ratio [38,39]. Nevertheless, the reflectance data of SiO is not just a simple superposition of those of Si and SiO2, but rather shows a combination of Si and SiO2 features that cannot be explained by the RM model [39]. On the other hand, Temkin suggested the microscopic mixture model, in which Si-like and SiO2-like regions (diameters: 5 to 10 Å) are mixed and their boundaries serve as the connectivity region [40]. By simply assuming the existence of Si–Si and Si–O bonds in the SiO structure, Temkin expected some Si atoms to be connected to an oxygen atom as the boundary region. Analytical methods such as infrared absorption (IR) spectroscopy, X-ray photoelectron spectroscopy (XPS), ultraviolet photoemission spectroscopy (UPS), and 29Si magic-angle spinning nuclear magnetic resonance (MAS NMR) were used to further investigate the structural model of SiO [4144]. Using the pair distribution function (PDF) and transmission electron microscopy (TEM), Schulmeister et al. reported the existence of an interfacial region outside the Si and SiO2 domains, and that the Si and O atoms are chemically inhomogeneous at the nanometer scale [45]. Hohl et al. suggested an interface cluster mixture model similar to the model of Schulmeister et al., indicating that 10% of the SiO connects the surrounding, very small, and atomically amorphous Si and SiO2 as an interface [46]. Recently, developments in analytical technology have enabled more precise and elaborate investigation of the SiO structure [47,48]. Using atom probe tomography (APT), the atomic Si domains were shown to be distributed non-uniformly in 3D atom maps of SiO powder, even though the SiO was not disproportionated [47]. Hirata et al. examined the SiO structure in detail using angstrom-beam electron diffraction (ABED) and TEM/scanning transmission electron microscopy (STEM) (Fig. 1) [48]. By investigating three areas (dark, bright, and interface areas) in the high-angle annular dark field scanning TEM (HAADF-STEM) images of SiO, they suggested the presence of Si-like and SiO2-like regions, with interfacial regions of suboxide (SiOx) that contained atomic coordinates ranging between Si-(3Si, O) and Si-(Si, 3O).
Much insight into the structure of SiO has been gained through continued research and development of analytical methods. Based on these structural investigations, the a-SiO structure could be identified as a mixture of angstrom-sized Si and SiO2 domains, together with interfacial regions where Si and SiO2 contact each other. Although most studies so far were confined to SiO, the results are expected to broaden our understanding of the SiOx structure.

ELECTROCHEMICAL PROPERTIES OF SiOx

The x value of SiOx

The value of x (0 < x < 2) affects the electrochemical performance of SiOx, as shown in Table 1 [4957]. The specific capacity and ICE of SiOx tend to decrease with increasing x. Considering the structural model of SiOx, a higher x value indicates an increase in the SiO2-like regions, which can change the intrinsic properties of the material. Al-Maghrabi et al. designed combinatorial SiOx thin films to identify the electrochemical property changes according to the oxygen ratio of SiOx [53]. Based on the voltage profiles and dQ/dV plots of these thin-film electrodes, they proposed three models for the electrochemical mechanism. In the model most consistent with the experimental data, Si reacts with lithium to produce Li3.75Si, and SiO2 reacts with lithium to produce Li4SiO4 as a Li-inactive phase, indicating that increasing the x value results in a higher irreversible capacity. Because those authors suspected that the formed Li4SiO4 phase enhances the diffusion of Li into Si and buffers volume expansion, they proposed that an optimal oxygen content could improve the electrochemical performance of the SiOx anode material. Haruta et al. used radio frequency (RF) magnetron sputtering to synthesize SiOx with x = 0.02, 0.21, 0.48, 1.09, and 1.78 and compared their electrochemical performance and morphology changes [50]. Their results revealed that a higher x increased the irreversible reaction of Li4SiO4 formation, which led to a drop in ICE but better cycle performance. Through the buffering role of the Li4SiO4 phase, cross-sectional SEM images of the cycled SiOx electrode showed less morphological changes and crack formation compared to the pure Si electrode. Cho et al. controlled the surface oxygen content of SiOx using RF magneton sputtering and investigated the resulting structural changes in SiOx thin films. X-ray photoelectron spectroscopy (XPS) revealed that the SiOx film underwent chemical changes during cycling, with the most significant changes observed at the highest oxygen content (x = 2). These results suggest that the solid electrolyte interphase (SEI) layer can be engineered by tuning the surface oxygen content of SiOx [58]. During ball milling, the value of x in SiOx can also be tuned by controlling the air exposure time or varying the mixing ratio of Si and SiOx [57,59,60]. These findings demonstrate that the electrochemical performance of SiOx varies significantly with the x value. Because of the trade-off relationships among ICE, specific capacity, and cycle performance, an appropriate x value for SiOx could be important for achieving the required electrochemical performance.

Disproportionation of SiOx

During disproportionation, amorphous SiOx separates into crystalline Si and SiO2 phases, which can affect the electrochemical performance [6164]. Disproportionation can be induced thermally by annealing SiOx powder with or without a catalyst (Fig. 2). In the XRD data of annealed SiO, Bragg peaks for crystalline Si phase were observed at temperatures above 800°C. In the presence of NaOH as a catalyst, both crystalline Si and SiO2 phases were observed in the XRD data of porous SiO at temperatures above 700°C. Another study reported the emergence of a crystalline SiO2 phase at above 1350°C without a catalyst [61]. Park et al. reported the characteristics and electrochemical behavior of SiO, whose degree of disproportionation was varied by heat treatment at different temperatures [35]. Based on the XRD, high-resolution TEM (HRTEM), and 29Si MAS NMR data of disproportionated SiO and pristine SiO, those authors proposed that the crystalline Si phase emerges at 800°C, whereas the crystalline SiO2 phase does not emerge until 1200°C. Upon increasing the heat-treatment temperature, the average crystallite size of Si increased from 5 nm (1000°C) to 30 nm (1200°C). Because of the uniformly dispersed crystalline Si particles and surrounding silicon suboxide matrix, electrodes with SiO heat-treated at 1000°C showed the highest ICE and capacity retention. However, electrodes with overly disproportionated SiO (1200°C) showed poor electrochemical properties, because the Li-inactive SiO2 phase hinders the alloying reaction of Si with Li. Such poor electrochemical performance may be related to the larger domain size of the SiO2 phase in the disproportionated SiO. It has been reported that high-energy mechanical milling (HEMM) can reduce the particle size of disproportionated SiO, re-activate Si, and improve the cycle performance compared to pristine SiO electrodes [65].
SiOx also can be disproportionated via electrochemical reactions with Li during cycling. Laser-assisted atom probe tomography (La-APT) was used to visualize Si domains in SiO before and after 419 cycles [66]. The results showed that the Si domains, which were randomly scattered in the SiO matrix before cycling, agglomerated and formed larger ones after cycling, implying that electrochemical disproportionation can occur because of the unstable structure of a-SiO.

Electrochemical reaction mechanism of SiOx in LIBs

The reaction mechanism between SiOx, especially SiO, and lithium is challenging to investigate because of (1) the uncertain amorphous structure of SiOx and (2) the difficulty in detecting the light Li atoms using electron beam analysis. Furthermore, because the electrochemical mechanism of SiOx is affected by its structural and compositional properties, such as the x value and degree of disproportionation, it has only been investigated in case studies (Table 2) [55,6679]. Nagao et al. reported a negative peak for lithiated SiO in the RDF pattern derived from neutron elastic scattering (NES), which is related to lithium oxide compounds apart from the LixSi alloy formed by the reaction of silicon with lithium [67]. Kim et al. reported that Li4SiO4 and Li2O were formed during the lithiation of SiO, based on 29Si-NMR and 7Li-NMR analyses [70]. The SiO2 in SiO reacts irreversibly with Li to form the Li4SiO4 phase, whereas Si reacts reversibly with Li to form the LixSi alloy phase. This two-phase reaction of SiO was also confirmed by dilatometric analysis. Kim et al. reported that SiO reacts with Li to form Li15Si4, Li-Si-O (especially Li4SiO4), and Li2O phases, based on the dQ/dV plot, 7Li MAS NMR and 29Si MAS NMR spectra, and HRTEM images [71]. From these results, they suggested that SiO undergoes a three-phase reaction following the revised random mixture model (a-Si, interphase boundary layer, and a-SiO2), in which the interphase boundary layer reacts with lithium to form the Li2O and Li15Si4 phases. Using a thermodynamic approach, Yasuda et al. calculated the reaction pathway for the lithiation of SiO and SiO2 with respect to the Li–Si–O terminal phase diagram [75]. After comparing the voltage window and capacity for each reaction stage, they determined that the possibly formed phases were Li4SiO4, Li2O, and Li13Si4. The Li22Si5 phase, considered to be a fully lithiated form of lithium silicon alloy, could not be formed thermodynamically; instead, Li13Si4 was the favorable phase for the lithiation of SiO. Recently, Kitada et al. reported phase changes in SiO during cycling through in-situ 7Li as well as ex-situ 7Li and 29Si MAS NMR techniques [79]. They compared pure Si, a-SiO, and disproportionated SiO (d-SiO) electrodes by marking the lithiation process of Si in three major steps 1–3, as shown in Fig. 3a. While the insitu 7Li MAS NMR data of the pure Si electrode was asymmetric during charging and discharging, that of the a-SiO electrode was symmetric (Fig. 3b). Their finding suggested that pure Si transformed into the c-Li15Si4 phase during the lithiation process (phase 3), but a-SiO did not form the c-Li15Si4 phase and instead formed a metallic m-LixSi phase (x = 3.44), which did not undergo a phase transition to the crystalline lithium silicide phase. This difference in electrochemical process between pure Si and a-SiO electrodes could explain the improved cycle performance of a-SiO, in that a-SiO does not suffer severe structural changes caused by c-Li15Si4 formation. Those authors also suggested that Li4SiO4 was produced during the initial lithiation reaction, and that Li2SiO3 was produced during the subsequent cycling of a-LixSi and Li4SiO4.
The internal Si in SiOx undergoes a reversible reaction with the lithium-silicon alloy, and the remaining SiOx or SiO2 regions react with lithium to form a Li-Si-O phase, which is generally accepted as Li4SiO4. However, the formation of lithium compounds and the value of x of LixSi in the fully lithiated SiO remain uncertain. For example, Li2SiO3 could emerge during the delithiation of Li4SiO4 or due to the electrolyte decomposition reaction [80]. These differences in the experimental data may be related to the different structures and compositions of SiOx materials used in each study.

STRATEGIES TO DEVELOP ADVANCED SiOx-BASED MATERIALS

Owing to the formation of a buffer phase, SiOx shows better cycle performance compared with other Si-based materials, but it has not yet met the commercialization requirements. Due to consumption of Li+ from the cathode due to irreversible formation of the Li-Si-O phase, SiOx displays a low ICE that lowers the energy density of LIBs. Also, the low electrical conductivity of SiOx reduces its reliability. Many efforts have been made to solve these two problems and enhance the electrochemical performance of SiOx anode materials.

SiOx-carbon composites

Carbon has good intrinsic electronic conductivity and forms electronically conductive networks to enable fast electron transport [81,82]. In SiOx-carbon composites, carbon coating has been used to increase the electrical conductivity and minimize the direct contact of active materials by forming a protective layer [55,82,83]. The coating processes are highly scalable and cost-effective, with the additional advantage of stabilizing the SEI layer [84]. Carbon coatings also reduce volume expansion by relieving stress owing to their ductility, sufficient porosity, and flexibility [85]. Another strength of carbon coating is the ease of application and flexibility compared to other coating methods [32,86]. Wet-chemistry synthesis using the pyrolysis of organic materials is considered one of the most useful, scalable, and easiest approaches. Chemical vapor deposition (CVD) and physical vapor deposition (PVD), which vaporize carbon or carbon precursors to deposit carbon on a substrate, are capable of forming uniform and stable coating layers. Research on carbon-coated Si electrodes showed that coating eliminated the parasitic reaction between the electrode and ethylene carbonate (EC)/propylene carbonate (PC), and the formed surface contained an electronic contact network that enhanced the electrochemical properties [87,88]. Due to its low intrinsic electrical conductivity, SiOx is especially suited for carbon coating to enhance its electrochemical performance [89,90]. In an early report in 2007, Kim et al. coated carbon on SiO via wet chemical pyrolysis using polyvinyl alcohol [32]. The coated electrode showed a capacity of 800 mAh g–1 and excellent cycle performance, which can be explained by the electronic contact bestowed on each particle through carbon coating that remained despite the volume change. Zhai et al. prepared polypyrrole (PPy) to prepare carbon-coated SiO composites via in-situ polymerization and high-temperature carbonization. The SiO/M@cPPy showed an ICE of 63.1% and an initial discharge capacity of 1657.9 mAh g–1 [91]. Using a similar approach, researchers doped the carbon layers with heteroatoms by using nitrogen-, fluorine-, or phosphorus-containing polymers via wet-chemistry pyrolysis, effectively enhancing the electrochemical properties [92,93]. You et al. synthesized Si/SiOx@F-C composites by in-situ pyrolysis of polyvinylidene fluoride (PVDF), which served as both carbon and fluorine source [92]. High-temperature decomposition of PVDF created a fluorine-doped carbon layer on the Si/SiOx surface. XPS analysis confirmed that the fluorine-doped coating significantly increased the LiF concentration in the SEI layer, contributing to enhanced interfacial compatibility. The Si/SiOx@F-C electrode showed an ICE of 79% and a high capacity retention of 82.12% after 2400 cycles.
Liu et al. used benzene as a carbon precursor to fabricate uniform carbon-coated SiO composites in a fluidized bed CVD reactor [86]. The carbon coating introduced a conductive network and pre-set voids, which compressed the polarization, enhanced the cycling performance, and reduced the irreversible electrode expansion. Simultaneously, secondary particles formed by the agglomeration of carbon-coated SiO particles acted as an additional volume buffer. The C-SiO composite electrode showed excellent capacity retention of 88% after 50 cycles. Yi et al. compared SiO anodes coated with pitch- and acetylene-derived carbon via CVD, and demonstrated that the acetylene-based coating featured a higher degree of graphitization and more uniform surface coverage [94]. These features contribute to improved electrical conductivity and structural integrity, leading to superior Li+ transport kinetics and cycling performance. Similarly, Zhang et al. reported on a dynamic CVD method that uses C2H2 to deposit a dense amorphous carbon layer on SiO, which improved both the electrochemical and mechanical performance by enhancing structural strength (Young’s modulus: 35 GPa) and suppressing volume expansion (9.3%) [95]. In 18650-type cells, C2H2@SiO achieved a high ICE of 87.58% and maintained 85.7% of its capacity after 1000 cycles at 25°C. In another study, the same researchers constructed a low-impedance carbon shell on SiOx via sequential gas-phase coating using C3H8 and C₂H₂ [96]. The resulting SiO@C3H8@C₂H₂ structure featured short-range vertically aligned and long-range lamellar carbon networks, which reduced the charge transfer resistance (Rct: 13.3 Ω) and contact resistance (Rc: 1.6 Ω) compared to pristine SiOx (Rct: 47.5 Ω, Rc: 7.2 Ω). This architecture showed a high ICE of 83.7% and achieved excellent capacity retention of 86.6% after 1000 cycles during full-cell evaluation at 0.5 C.
Exploiting the aforementioned good intrinsic properties of carbon materials, various SiOx-carbon composite structures have been actively studied, such as yolk-shell, sugar apple-shaped, and nanowire structures [90,97,98]. Among them, composites formed from organosilica materials have the strength of naturally coexisting silica and carbon coating precursors, while being eco-friendly, easily scalable, and do not require expensive reagents [99101]. For example, rice husks (RH) contain both carbon precursors (such as lignin, cellulose, and hemicellulose) and Si-based inorganic compounds [101103]. Cui et al. used RH to fabricate porous SiOx@C composites via carbonization and aluminothermic reduction [102]. The microporous SiOx@C composite electrode exhibited a capacity of 1230 mAh g–1 (0.01–3 V vs. Li/Li+) and excellent cycle life with less than 0.5% capacity loss after 200 cycles. Recently, Zhu et al. fabricated atomic-level distributed carbon in a SiOx matrix using an organoalkoxysilane precursor [104]. While existing wetchemistry or physical mixing methods result in an inhomogeneous carbon distribution, their approach produced a homogeneous SiOx-carbon composite. In organosilica nanospheres created by organic-inorganic hybrid composition, the organic alkly group of phenylene-bridged mesoporous organosilicas (PBMOs) is homogeneously distributed, and a homogeneous carbon matrix can be achieved through carbonization (Fig. 4a). A uniform carbon material was formed by graphitization of aromatic rings, as revealed by Raman spectroscopy and HAADF-STEM with energy-dispersive X-ray spectroscopy (EDS) (Fig. 4b,c). In the synthesized carbon composite with homogeneous atomic distribution, the molecularly interconnected carbon facilitated electron transport, and the micropores increased the Li+ diffusion rate, resulting in a high initial discharge capacity of 765 mAh g–1 and a superior cycle capacity of 501 mAh g–1 after 300 cycles.
In addition to conventional SiOx-carbon composites, Si/C hybrid structures have been developed to combine the mechanical robustness and electrical conductivity of carbon materials with the high theoretical capacity of silicon-based materials [105108]. These hybrid designs incorporate graphite, graphene, or engineered carbon shells to enhance the interfacial stability and ensure long-term cycling performance. Wang et al. introduced a dual interfacial reinforcement strategy employing graphene and carbon coatings to simultaneously enhance the electronic/ionic transport and structural stability during extended cycling [105]. The Gr-SiOx@C electrode retained a discharge capacity of 652 mAh g–1 over 1000 cycles at 1 C with a capacity retention of 97%. Xiao et al. developed a spherical natural graphite(SNG)/hollow-SiOx@C composite by integrating hollow-structured SiOx with natural graphite and a pitch-derived carbon coating [106]. The pitch functioned as both a conductive matrix and a binder, effectively maintaining cohesion between the SiOx and graphite particles. This composite achieved an initial capacity of 465mAh g–1 and maintained excellent cycling stability with a capacity retention of 93% over 500 cycles at 0.5 C.

Heteroatom doping

Heteroatom doping adds a specific element to the SiOx matrix through two types of approaches. In the first one, the substance is added to SiOx to form a simple composite with improved electrochemical performance due to the intrinsic characteristics of the additive, such as high conductivity and mechanical stability [109118]. In the second approach, the ICE and cycle performance are enhanced through reaction of the heteroatom (M) to form lithium-inactive phases before cycling [119122].
A major advantage of composite materials is their ability to combine characteristics from different components to match the requirements for diverse applications. Ball milling is a simple process that can create a wide variety of composites. Because heteroatom doping is intended to compensate for the disadvantages of SiOx, the heteroatom sources tend to have high electrical conductivities and stable structures to maintain an electric network with buffering roles. Tin (Sn) can undergo Si-like Li alloy reaction and shows good ductility and high electronic conductivity [109]. A SiOx/Sn/C hybrid composite was formed by adding Sn particles to a 2D SiOx sheet, followed by carbon coating to improve the electronic conductivity, tap density, and inner structural stability. The hybrid composite electrode showed an improved cycle performance of 838.5 mAh g–1 after 100 cycles, and the introduction of Sn enhanced the tap density to 0.75 g cm–3. Gu et al. fabricated a SiOx@SnO2@C composite by utilizing a lithiated SnO2 material with high conductivity and excellent structural stability, resulting in suppressed Si pulverization and improved cycle performance of the electrode [110]. Fu et al. fabricated a Sn/SiO composite by ball milling nanoscale Sn and SiO together [123]. The re-activation of Li+ in lithium irreversible phase Li2O was possible owing to the high electrical conductivity of Sn particles. Therefore, the ICE of the Sn/SiO composite electrode was higher than that of the pristine SiO electrode (85.5% vs. 66.5%). Copper (Cu) also has a high electrical conductivity (5.8 × 107 S cm–1) and good ductility for use as a dopant in SiOx anode materials [112]. A SiO/Cu/expanded graphite (EG) composite was synthesized to exploit the buffering role of EG and Cu as conductive additives. The composite electrode had a high capacity and cycle performance of 836 mAh g–1 at the 100th cycle, because the Cu network and EG had excellent ductility and maintained facile electrical conduction while effectively alleviating stress in the electrode. TiO2 has also been used as a dopant to improve the electrochemical performance of SiOx [113117]. Using the sol-gel method, Li et al. trapped TiO2 nanoparticles in a SiOx matrix and then applied carbon coating to create a watermelon-like SiOx–TiO2–C composite (Fig. 5a) [113]. Anatase TiO2 exhibits high Li+ conductivity, excellent electrical conductivity in the lithiated form, and fast Li+ insertion/extraction through its (001) plane. The prepared SiOx–TiO2–C composite electrode showed a great cycle capacity of 910 mAh g–1 after 200 cycles (Fig. 5b). Bae et al. coated Si/SiOx with black TiO2-x, and the corresponding electrode showed a high reversible capacity of 1200 mAh g–1 with excellent performance up to 100 cycles [117]. TiO2-based materials also have the unique ability to improve thermal reliability [116]. The discharge capacity of black TiO2-x-coated Si/SiOx electrode was improved by 3%, and its capacity recovery rate was 15% higher than that of the pristine Si/SiOx electrode after staying in the charged state at 60°C for 48 h.
Some dopants enhance the electrochemical performance by modulating the SEI through the formation of stable lithium compounds or reversible conversion reactions. Zhao et al. fabricated a Se-doped C/SiOx composite via solid-phase sublimation, in which Se was deposited on both the interior and exterior pore walls [124]. Se doping promoted the formation of Li₂Se in the SEI, thereby suppressing Li₂O formation and reducing the interfacial resistance. The Se@C/SiOx electrode achieved a high ICE of 82.4% with a discharge capacity of 844.4 mAh g–1 at 0.1 C. Similarly, Ling et al. synthesized a Co-doped SiOx composite by coating SiO₂ spheres with a Co2+-doped polymer, followed by carbonization. The embedded Co nanoparticles enhanced charge transfer kinetics and contributed to additional reversible capacity through the conversion of Li₂O [125]. The SiOx@NC-Co electrode showed an ICE of 62.2% with a discharge capacity of 890.95 mAh g–1 at 0.1 A g–1.
After the chemical reaction of SiOx with the heteroatom element M, the resultant MOx or M–Si–O Li-inactive phases may increases ICE. These phases could further improve the electrochemical properties through their own characteristics such as good mechanical strength. Zhang et al. fabricated a C–SiO–MgSiO3–Si anode material with high ICE by reacting MgO with SiO to form MgSiO3 [120]. The Li-inactive MgSiO3 buffer phase and the mesoporous secondary particles improved the ICE and cycle performance of the C–SiO–MgSiO3–Si secondary particle electrode, with a high ICE of 78.3% and a great capacity of 1608 mAh g–1. Similarly, Youn et al. prepared a Si/Mg2SiO4 composite material via the MgH2-driven magnesiation of SiO, with Mg2SiO4 acting as a Li-inactive buffer phase [126]. The endothermic dehydrogenation of MgH2 effectively suppressed coarsening of the Si domains. As a result, the Si/Mg2SiO4 nanocomposite electrode showed a high ICE of 89.5% and an initial discharge capacity of 1108 mAh g–1. Using a different doping element, the same group demonstrated a porous Si/Ca–Si–O nanocomposite via the calciothermic reduction of SiO using CaH2, in which the Li-inactive Ca–Si–O domains served as the buffer phase [127]. Subsequent HEMM resulted in a porous microstructure capable of accommodating volume changes in Si during cycling. This porous Si/Ca–Si–O composite electrode showed a high ICE of 86.4% and maintained a discharge capacity of 55% after 200 cycles. Alumina also has sufficiently high structural stability to suppress volume expansion during the LixSi alloying reaction, and it can compensate the low electrical conductivity of SiOx [128,129]. Kim et al. prepared a carbon-coated Si–SiOx–Al2O3 composite via a two-step process: (1) aluminothermic reduction of SiO2 to prepare a Si-embedded SiOx–Al2O3 composite, and (2) carbon coating using naphthalene as a carbon precursor. The carbon-coated Si–SiOx–Al2O3 composite electrode showed a high ICE of 74.3% and excellent cycle performance of 500 mAh g–1 after 500 cycles [122].

Prelithiation

Prelithiation is under active investigation as a way to increase the ICE of SiOx electrodes through heteroatom doping. The goal is to reduce active lithium loss (ALL) due to the initial irreversible formation of lithium oxide and Li–Si–O in SiOx, which decreases the ICE [130]. Because a full cell contains a limited amount of Li+, this irreversible reaction causes internal capacity loss and a lower energy density [130]. Prelithiation works by inserting lithium before cycling. Unlike other heteroatom doping methods, prelithiation artificially creates phases that are also formed during the actual electrochemical reactions of cells, such as lithium silicates, lithium oxide, and even SEI layers. Besides an electrochemically irreversible phase, it can also create a reversible LixSi alloy phase that enables the ICE to even exceed 100% [131].
We divide prelithiation methods into five types: (1) contact, (2) impregnation, (3) short circuiting, (4) vapor deposition, and (5) thermal. In the contact method, Li metal is directly attached to the SiOx anode. This method is scalable, easily processable, and allows the formation of an SEI layer on the electrode. Using the contact method, researchers reported improved performance of SiOx full cells and lithium-ion capacitors [132,133]. However, its utilization remains difficult due to poor reaction control, possible Li-dendrite formation, inhomogeneous solid-solid reaction, and safety issues associated with Li metal. Recently, a homogeneous and controllable contact method was proposed to prelithiate SiOx electrodes by using polyvinyl butyral-coated carbon nanotube as a resistance buffer layer (RBL) [134]. The RBL inhibits direct contact between the Li metal foil and the SiOx anode. By making contact prelithiation homogeneous and controllable, RBL-regulated prelithiation improves the ICE from 79% to 89% and shows stable cycle performance up to 200 cycles in full-cell applications, compared to the electrode without prelithiation.
Impregnation or organochemical prelithiation utilizes the characteristics of Li metal melted with benzene-based substances in organic solvents. In this method, electrodes or anode materials are prelithiated by immersion in lithium-containing benzene-organic solutions [135137]. The impregnation approach is scalable and could achieve relatively uniform reaction. Tabuchi et al. performed organochemical prelithiation by immersing electrodes in a Li-organic complex solution consisting of Li metal and naphthalene in butyl methyl ether [135]. After 72 h, the prelithiated SiO electrode showed a discharge capacity of 670 mAh g–1 without charging. Huang et al. used a prelithiation agent in an SLMP-SBR-toluene (SST) suspension, and the ICE was improved from 66% to 70%–120% [138]. Li et al. immersed negative electrodes in a LAC (N-LAC and B-LAC) solution for a controlled period and at specific temperatures [139]. For N-LAC solution at 60 min, the cells showed an ICE of 93.5%. The cell prelithiated with N-LAC also had a longer cycle life and a higher energy density. In another study, the SiO electrode was prelithiated by immersion in a solution of lithium-2-fluorobiphenyl-2-methyltetrahydrofuran (Li-FBP-MeTHF) [140]. The SiO electrode prelithiated with a LiF oligomer layer exhibited high hydrophobicity and maintained good air stability. Its ICE also increased to 101.7%.
In the short-circuit prelithiation method, artificial external short circuits are created between the Li source and the electrode. This method is both scalable and controllable [131,141]. Kim et al. reported a roll-to-roll short-circuit prelithiation process using Li metal foil, as shown in Fig. 6a,b [131]. The penetration speed and degree of reaction were controlled by changing the resistance applied to the external circuit. XPS and time-of-flight secondary ion mass spectrometry (TOF-SIMS) analyses confirmed that the artificially formed SEI layer was identical to that formed during the actual electrochemical process. The longer the prelithiation time, the higher the ICE (up to 107.9%), but the optimal efficiency was probably only 94.9% owing to Li dendrite formation caused by overcharging, which degraded the electrochemical properties (Fig. 6c,d).
In vapor deposition, a Li metal source is vaporized to induce a gas-phase reaction on the electrode. Takezawa et al. vaporized metallic Li using a thermal evaporator to manufacture prelithiated a-SiOx electrodes with different x values (0.51, 0.68, and 1.02), by adjusting the required amount of Li metal according to the oxygen ratio [142]. The prelithiated SiO1.02 electrode showed a high ICE of 118% and a capacity of 1041 mAh g–1.
In the thermal method, heat treatment is applied to induce chemical reactions between SiOx and Li source, thereby directly prelithiating the active material [74,143145]. Zhao et al. prepared LixSi nanodomains in a Li2O matrix through the reaction between SiO and Li metal [144]. The resultant electrode showed an excellent capacity of 1240 mAh g–1 without charging, even after exposure to a humidity of 40% RH for 6 h. Yom et al. demonstrated a prelithiated SiO electrode with 82.1% ICE by producing Li2SiO3 and Li4SiO4 as irreversible phases through heat treatment of Li powder and SiOx [74]. In the Li metal-free prelithiation method proposed by Chung et al., LiH is dehydrated at elevated temperatures above 470°C [146,147], and the released Li acts as a precursor for the preemptive formation of lithium silicates (the primary cause of low ICE for SiO) [146]. Prelithiation of SiO induces a phase transformation into Si and Li₂SiO₃ with an optimized topological arrangement [147]. The material treated at 650°C exhibited ~90% ICE and stable cycling over 300 cycles. This improvement in performance is due to the uniform nanoscale distribution of the active and inactive phases, which significantly reduces volume expansion.

Structural modification

Structural modification aims to improve the electrochemical performance by morphological changes, such as pore generation and nanostructuring, in order to control volume expansion and enlarge the surface area to reduce the Li diffusion length and improve the kinetics. In 2007, Kim et al. adjusted the particle size of SiOx through a ball-milling process to increase the mechanical stability and significantly improve the cycle performance [32]. The ball-milled SiO–C composite electrode showed an excellent cycle life because of the homogeneous volume expansion, owing to the uniform particle size and shape with free voids created by the carbon coating. Intensive research efforts have been devoted to enhancing the electrochemical properties of SiOx through structural modifications by creating ball-milled SiOx–C composites [86,148150].
A porous structure facilitates the transport of Li+ through a larger surface area and improves the cycle life through pre-set voids [25,151]. Park et al. prepared a c-Si-embedded Si/SiOx complex via the hydrothermal reduction of hydrogen silsesquioxane (HSQ) prepared by the sol–gel process involving triethoxysilane (Fig. 7a) [22]. This non-toxic and highly scalable sol-gel process produced a nanosized porous structure with superior expansion control. As a result, the electrode showed 78% capacity retention after 100 cycles and 20.8% volume expansion after two cycles (Fig. 7b,c). Similarly, Lee et al. formed porous SiO particles through HF etching of Ag-deposited SiO, which was formed by the reaction between AgNO3 and SiO. The porous SiO electrode delivered a high specific capacity of 1520 mAh g–1 with stable capacity retention of 1490 mAh g–1 after 50 cycles [152]. Park et al. obtained a porous SiO composite by mechanochemically oxidizing an inexpensive Si powder while simultaneously reducing ZnO powder using a high-energy ballmilling process [153]. The porous SiO composites exhibited enhanced electrochemical performance compared to pristine SiO, which was attributed to the microstructural changes. Raza et al. created double-layered SiOx/Mg2SiO4/SiOx composites through the magnesiothermic reduction of micro-sized SiO with Mg metal powder [154]. The composite electrode showed an ICE of 75.6% and a capacity retention 86% after 300 cycles. Core-shell matrices were also studied to utilize the characteristics of both the core and shell components. For example, core-shell SiOx synthesized by gradually increasing the x value of SiOx from the inside to the outside combines both the high CE and high specific capacity of Si and the superior cycle reliability of SiOx [155157].
Nanostructuring enhances the electrochemical performance of SiOx by controlling its structure at the nanoscale [158,159]. For example, controlling the direction of SiOx expansion can reduce crack formation during the volume expansion of Si, increase electronic conductivity by forming a conductive network, and decrease the Li diffusion length, resulting in superior Coulombic efficiency, cycle performance, and rate capability. Reported SiOx nanostructures include nanorods [34], nanotubes [160], nanocoils [161], nanowires [162], and 3D gyroid structures [163]. Li et al. reported core–shell porous carbon-coated SiOx nanowires (pC-SiOx NWs) through the self-sacrificial SiOx nanowire formation using bimodal mesoporous silica (BMS) [90]. The 1D structure effectively alleviated the strain, maintained the structure, and prevented pulverization during repeated cycles. The corresponding electrode demonstrated excellent cycle performance: the capacity did not decrease from the 2nd to the 150th cycle and remained at 1060 mAh g–1 after 100 cycles.
While nanosized particle modification can create electrodes with excellent cycle performance and rate capability, excessive and irreversible SEI formation may occur because of the increased surface area. The trade-off between nanosized modification and excessive SEI formation should therefore be carefully optimized. Yu et al. proposed a hybrid structure comprising active SiO particles and carbon nanofibers (SiO/CNFs) [164]. The nanofibrous carbon skeleton prevented particle agglomeration and restricted the volume expansion of SiO. A capacity retention of 73.9% was achieved after 400 cycles, because the structure not only provided a conductive network for active SiO nanoparticles but also surmounted agglomeration sufficiently.

CONCLUDING REMARKS

SiOx has been actively studied as an anode material for LIBs, due to its better specific capacity compared to graphite and better cycle life compared to Si. This mini-review discusses studies on the structure and electrochemical characteristics of SiOx. While the unique properties of SiOx hinder clear identification of its structure and electrochemical mechanism, this material can be considered as having Si and SiO2 arranged on an atomic scale with interfacial regions. It shows superior cycle life compared to Si because of buffer phases introduced by lithiation, such as lithium silicates and lithium oxide. However, the adoption of SiOx as the sole anode material for commercial LIBs is hampered by a low ICE due to irreversible phase formation reaction and insufficient long-term cyclability compared to graphite. We discussed some promising efforts to overcome the weak electrochemical performance of SiOx through four approaches: (1) SiOx–carbon composites, (2) heteroatom doping, (3) prelithiation, and (4) structural modification.
Despite recently reported enhancements in SiOx-based anode materials, intensive research is still needed before SiOx can be used by itself as the anode material for commercial LIBs. In terms of reliability, full cells employing SiOx-based electrodes should maintain their performance for more than 500 cycles. For the emerging EV industry, thermal stability, thermal safety, and high-rate capability are also important parameters. Moreover, the x value of SiOx should be optimized according to the target performance (high C-rate, capacity, ICE, and cyclability) before adopting additional approaches to enhance the electrochemical performance of SiOx-based materials. By applying various techniques (including the simple application of carbon coatings) to enhance the properties of SiOx and conducting multidisciplinary research, we believe that SiOx-based materials have great application potential as core components for LIBs with high energy density.

Fig. 1.
(a) Reconstructed heterostructure model of amorphous SiO. The inner part corresponds to an amorphous Si cluster, and the outer part is an amorphous SiO matrix. Blue, red, and green circles denote Si and O in amorphous 2 SiO2 and Si in the Si cluster, respectively. (b) Experimental and simulated X-ray total structure factor S(Q) curves. (c) Fractions of the five atomic coordinates found in amorphous SiO. (d) Normalized intensity profiles constructed by integrating ABED patterns obtained from the dark, bright, and interface areas in HAADF-STEM images. Reproduced with permission from Ref. [48]. © 2016 Springer Nature.
jecst-2025-00773f1.jpg
Fig. 2.
(a) XRD patterns of pristine and disproportionated SiO samples. Reproduced with permission from Ref. [64]. © 2022 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. (b,c) XRD patterns showing the disproportionation of SiO (b) annealed at different temperatures for a fixed time of 1 h and (c) annealed for different times at a fixed temperature of 800°C. Reproduced with permission from Ref. [151]. © 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
jecst-2025-00773f2.jpg
Fig. 3.
(a) Schematic structures of LixSi phases at different Li:Si ratios (x) formed by electrochemical lithiation of pure Si, and a comparison with motifs found in crystalline phases at similar compositions. Pink, green, and blue shaded areas represent three LixSi phases (1)−(3) that contain extended Si clusters, small Si clusters, and isolated Si atoms as found in c-Li15Si4, respectively. (b) In-situ 7Li NMR spectra and cell voltage of pure-Si and a-SiO measured at different times. The NMR intensity is color-coded from 0% (blue) to 100% (red). Peaks in the Super-P/CMC-Na matrix are labeled according to the environments of LixSi and Li as (i)–(vi): (i) extended Si clusters, phase 1; (ii) Li in Super-P/amorphous carbon; (iii) small Si clusters, phase 2; (iv) m-LixSi; (v) Li with nearby small Si clusters formed during conversion from crystalline Si to a-LixSi; (vi) c-Li15Si4-like structures, phase 3. Reproduced with permission from Ref. [79]. © 2019 American Chemical Society.
jecst-2025-00773f3.jpg
Fig. 4.
(a) Preparation of the porous silicon-based composite consisting of a homogeneous atomic-scale distribution of carbon (ASD-SiOC) nanocomposite. (b) Raman spectra of the PBMOs and the ASD-SiOC nanocomposite. (c) Dark-field TEM image of one ASD-SiOC nanocomposite particle and corresponding elemental mapping. Reproduced with permission from Ref. [104]. © 2019 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
jecst-2025-00773f4.jpg
Fig. 5.
(a) Schematic illustration for the first lithiation process of the SiOx-TiO2@C nanoparticle with a watermelon-like structure. (b) Cycling performance of SiOx-TiO2@C and SiOx@C electrodes at 1 A g–1. Reproduced with permission from Ref. [113]. © 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
jecst-2025-00773f5.jpg
Fig. 6.
(a) Schematic illustration of the prelithiation process of carbon-coated SiOx (c-SiOx) electrode and (b) its scalable roll-to-roll process scheme. (c) The first cycle voltage profiles of c-SiOx with different prelithiation durations. (d) Comparison of specific capacity and ICE after different prelithiation times. Reproduced with permission from Ref. [131]. © 2016 American Chemical Society.
jecst-2025-00773f6.jpg
Fig. 7.
(a) Schematic illustration of the C-coated Si/SiOx nanospheres. (b) Cycle performance and Coulombic efficiencies (inset) of C-coated Si/SiOx electrodes at a constant current of 0.2 C (200 mA g–1) for 100 cycles. (c) Cross-sectional FESEM images of C-coated Si/SiOx electrodes (left: pristine electrode, right: electrode after two cycles). Reproduced with permission from Ref. [22]. © 2014 American Chemical Society.
jecst-2025-00773f7.jpg
Table 1.
Reported electrochemical properties of SiOx with different x values.
x value 1st discharge capacity/mAh·g–1 1st charge capacity to 1.5 V/mAh·g–1 1st cycle efficiency/% Synthesis process Disproportionation Ref
0.02 3630 3370 92.9 RF magnetron sputtering X [50]
0.17 3448 3242 94.0 Vacuum electron beam evaporation X [55]
0.21 3440 3060 88.9 RF magnetron sputtering X [50]
0.40 3053 2502 82.0 Magnetron sputtering X [51]
0.48 2950 2700 91.5 RF magnetron sputtering X [50]
0.5 2630 2157 82 Ball milling X [56]
0.51 3052 2557 83.8 Vacuum electron beam evaporation X [55]
0.61 2526 2044 80.9 Magesiothermic reduction X [57]
0.68 2759 2064 75.0 Vacuum electron beam evaporation X [52]
0.70 2758 1993 72.3 Magnetron sputtering X [51]
0.71 2306 1806 78.3 Magesiothermic reduction X [57]
0.81 1881 1374 73.0 Magesiothermic reduction X [57]
0.83 1923 1500 78 Thermal plasma O [49]
0.95 1714 1150 67.1 Magesiothermic reduction X [57]
1.02 2418 1306 54.0 Vacuum electron beam evaporation X [55]
1.06 1595 1196 75 Thermal plasma O [49]
1.09 2300 1690 73.5 RF magnetron sputtering X [50]
1.10 2432 1433 58.9 Magnetron sputtering X [51]
1.18 701 Thermal evaporation O [54]
1.21 1485 1054 71 Thermal plasma O [49]
1.28 775 Thermal evaporation O [54]
1.34 2198 867 39.4 Vacuum electron beam evaporation X [55]
Table 2.
Phases observed in the electrochemical reaction of SiOx, together with the characterization methods used.
SiOx Lithiated Delithiated Analysis methods Ref
Amorphous, x = 0.94 Li3.5Si, Li4.4Si, LiOH, Li2O NES [67]
Li4SiO4, Li2SiO3
Amorphous LixSi, Li4SiO4, Li2Si2O5, Li2SiO3, Li2O Li4SiO4, Li2Si2O5, Li2Si2O3, Li2O, LixSi, SiOx XRD, XPS [68]
C-coated, Disproportionated LixSi, Li2SiO3, Li4SiO4 Li2SiO3, Li4SiO4 XRD, TEM [69]
Amorphous a-LixSi, Li2O, Li4SiO4 Li4SiO4 7Li NMR, 29Si NMR, TEM [70]
Amorphous Li15Si4, Li4SiO4, Li2O Li4SiO4, Li2O 29Si-MAS-NMR, 7Li-MAS-NMR, TEM, SAED [71]
C-coated, amorphous LixSi, Li2O, Li4SiO4 HRTEM [72]
C-coated Li3.75Si, Li4SiO4, Li2SiO3, SiO2 Si, SiO2, Li4SiO4, Li2SiO3 XANES, STEM, EELS, EDX, 7Li NMR, 29Si NMR, Molecular dynamics [73]
C-coated, Disproportionated, x = 1.04 Li15Si4 XRD [81]
Amorphous Li4Si, Li2O, Li4SiO4, Li2SiO3 LixSi, Li2O, Li4SiO4, Li2SiO3 XPS [55]
Amorphous Li2SiO3, Li4SiO4 XRD, TEM, SAED [74]
Amorphous, Disproportionated Li13Si4, Li4SiO4, Li2O Si, Li4SiO4 Thermodynamic calculation [75]
Amorphous Li21Si8, Li4SiO4 Li21Si8 HRTEM, XRD [76]
C-coated, Disproportionated LixSi, Li4SiO4 Li4SiO4 7Li-MAS-NMR, 29Si-MAS-NMR [66]
Amorphous a-LixSi, Li4SiO4 Li4SiO4 In-situ 7Li NMR, Ex-situ 29Si MAS NMR [77]
Amorphous, x = 1.58 Li6Si2O7, Li4SiO4, Li2Si2O5 Li6Si2O7, Li4SiO4 XPS, Grazing incidence XRD [78]
Amorphous, Disproportionated Li4SiO4, a-LixSi Li4SiO4, Si, Li2SiO3 In-situ 7Li NMR Ex-situ 7Li and 29Si MAS NMR [79]

REFERENCES

[1] J. B. Goodenough and K.-S. Park, J. Am. Chem. Soc., 2013, 135(4), 1167–1176.
crossref
[2] B. Scrosati and J. Garche, J. Power Sources, 2010, 195(9), 2419–2430.
crossref
[3] V. Etacheri, R. Marom, R. Elazari, G. Salitra and D. Aurbach, Energy Environ. Sci., 2011, 4(9), 3243.
crossref
[4] R. Marom, S. F. Amalraj, N. Leifer, D. Jacob and D. Aurbach, J. Mater. Chem., 2011, 21(17), 9938.
crossref
[5] T.-H. Kim, J.-S. Park, S. K. Chang, S. Choi, J. H. Ryu and H.-K. Song, Adv. Energy Mater., 2012, 2(7), 860–872.
crossref
[6] G. Jeong, Y.-U. Kim, H. Kim, Y.-J. Kim and H.-J. Sohn, Energy Environ. Sci., 2011, 4(6), 1986.
crossref
[7] H. Kim, G. Jeong, Y.-U. Kim, J.-H. Kim, C.-M. Park and H.-J. Sohn, Chem. Soc. Rev., 2013, 42(23), 9011.
crossref
[8] H. Kim, E.-J. Lee and Y.-K. Sun, Mater. Today, 2014, 17(6), 285–297.
crossref
[9] Y. Nishi, J. Power Sources, 2001, 100(1–2), 101–106.
crossref
[10] S. Basu, C. Zeller, P. J. Flanders, C. D. Fuerst, W. D. Johnson and J. E. Fischer, Mater. Sci. Eng., 1979, 38(3), 275–283.
crossref
[11] M. Armand and J.-M. Tarascon, Nature, 2008, 451, 652–657.
crossref pdf
[12] B. Nykvist and M. Nilsson, Nat. Clim. Change, 2015, 5, 329–332.
crossref pdf
[13] B. Dunn, H. Kamath and J.-M. Tarascon, Science, 2011, 334(6058), 928–935.
crossref
[14] J. K. Lee, C. Oh, N. Kim, J.-Y. Hwang and Y.-K. Sun, J. Mater. Chem. A, 2016, 4, 5366–5384.
crossref
[15] X. Su, Q. Wu, J. Li, X. Xiao, A. Lott, W. Lu, B. W. Sheldon and J. Wu, Adv. Energy Mater., 2014, 4(1), 1300882.

[16] B. A. Boukamp, G. C. Lesh and R. A. Huggins, J. Electrochem. Soc., 1981, 128, 725–729.
crossref pdf
[17] L. Y. Beaulieu, K. W. Eberman, R. L. Turner, L. J. Krause and J. R. Dahn, Electrochem. Solid-State Lett., 2001, 4, A137.
crossref
[18] M. N. Obrovac and L. Christensen, Electrochem. Solid-State Lett., 2004, 7, A93.
crossref
[19] S. C. Jung, J. W. Choi and Y.-K. Han, Nano Lett., 2012, 12, 5342–5347.
crossref
[20] R. Yi, F. Dai, M. L. Gordin, S. Chen and D. Wang, Adv. Energy Mater., 2013, 3, 295–300.
crossref
[21] S.-S. Suh, W. Y. Yoon, C.-G. Lee, S.-U. Kwon, J.-H. Kim, Y. Matulevich, Y.-U. Kim, Y. Park, C.-U. Jeong, Y.-Y. Chan and S.-H. Kang, J. Electrochem. Soc., 2013, 160(4), A751–A755.
crossref
[22] M.-S. Park, E. Park, J. Lee, G. Jeong, K. J. Kim, J. H. Kim, Y.-J. Kim and H. Kim, ACS Appl. Mater. Interfaces, 2014, 6(12), 9608–9613.
crossref
[23] S. Goriparti, E. Miele, R. P. Zaccaria and C. Capiglia, J. Power Sources, 2014, 257, 421–443.
crossref
[24] M. Jiao, Y. Wang, C. Ye, C. Wang, W. Zhang and C. Liang, J. Alloys Compd., 2020, 842, 155774.
crossref
[25] T. Chen, J. Wu, Q. Zhang and X. Su, J. Power Sources, 2017, 363, 126–144.
crossref
[26] Z. Liu, Q. Yu, Y. Zhao, R. He, M. Xu, S. Feng, S. Li, L. Zhou and L. Mai, Chem. Soc. Rev., 2019, 48, 285–309.
crossref
[27] S. Komaba, K. Shimomura, N. Yabuuchi, T. Ozeki, H. Yui and K. Konno, J. Phys. Chem. C, 2011, 115(27), 13487–13495.
crossref
[28] C. Huang, A. Kim, D. J. Chung, E. Park, N. P. Young, K. Jurkschat, H. Kim and P. S. Grant, ACS Appl. Mater. Interfaces, 2018, 10(18), 15624–15633.
crossref
[29] J.-Y. Li, Q. Xu, G. Li, Y.-X. Yin, L.-J. Wan and Y.-G. Guo, Mater. Chem. Front., 2017, 1, 1691–1708.
crossref
[30] J. W. Choi and D. Aurbach, Nat. Rev. Mater., 2016, 1, 16013.

[31] Y. Reynier, C. Vincens, C. Leys, B. Amestoy, E. Mayousse, B. Chavillon, L. Blanc, E. Gutel, W. Porcher, T. Hirose and C. Matsui, J. Power Sources, 2020, 450, 227699.
crossref
[32] J.-H. Kim, H.-J. Sohn, H. Kim, G. Jeong and W. Choi, J. Power Sources, 2007, 170, 456–459.
crossref
[33] S. Schweidler, A. Schiele, P. Hartmann, T. Brezesinski and J. Janek, J. Phys. Chem. C, 2018, 122(16), 8829–8835.
crossref
[34] Y. Ren and M. Li, J. Power Sources, 2016, 306, 459–466.
crossref
[35] C.-M. Park, W. Choi, Y. Hwa, J.-H. Kim, G. Jeong and H.-J. Sohn, J. Mater. Chem., 2010, 20(23), 4854.
crossref
[36] G. W. Brady, J. Phys. Chem., 1959, 63(7), 1119–1120.
crossref
[37] J. A. Yasaitis and R. Kaplow, J. Appl. Phys., 1972, 43(3), 995–1000.
crossref pdf
[38] H. R. Philipp and J. Non-Cryst, Solids, 1972, 8–10, 627–632.

[39] H. R. Philipp, J. Phys. Chem. Solids, 1971, 32, 1935–1945.
crossref
[40] R. J. Temkin, J. Non-Cryst. Solids, 1975, 17(2), 215–230.
crossref
[41] W. Y. Ching, Phys. Rev. B, 1982, 26, 6610–6621.
crossref
[42] R. Dupree, D. Holland and D. S. Williams, Philos. Mag. B, 1984, 50, L13–L18.
crossref
[43] M. Nakamura, Y. Mochizuki, K. Usami, Y. Itoh and T. Nozaki, Solid State Commun., 1984, 50, 1079–1081.
crossref
[44] F. G. Bell and L. Ley, Phys. Rev. B, 1988, 37, 8383–8393.
crossref
[45] K. Schulmeister and W. Mader, J. Non-Cryst. Solids, 2003, 320(1–3), 143–150.
crossref
[46] A. Hohl, T. Wieder, P. A. van Aken, T. E. Weirich, G. Denninger, M. Vidal, S. Oswald and C. Deneke, J. Mayer, and H. Fuess, J. Non-Cryst. Solids, 2003, 320(1–3), 255–280.
crossref
[47] H. Sepehri-Amin, T. Ohkubo, M. Kodzuka, H. Yamamura, T. Saito, H. Iba and K. Hono, Scr. Mater., 2013, 69, 92–95.
crossref
[48] A. Hirata, S. Kohara, T. Asada, M. Arao, C. Yogi, H. Imai, Y. Tan, T. Fujita and M. Chen, Nat. Commun., 2016, 7, 11591.

[49] S.-S. Suh, W. Y. Yoon, D.-H. Kim, S.-U. Kwon, J.-H. Kim, Y.-U. Kim, C.-U. Jeong, Y.-Y. Chan, S.-H. Kang and J.-K. Lee, Electrochim. Acta, 2014, 148, 111–117.
crossref
[50] M. Haruta, T. Doi and M. Inaba, J. Electrochem. Soc., 2019, 166(2), A258–A263.
crossref
[51] X. Meng, H. Huo, Z. Cui, X. Guo and S. Dong, Electrochim. Acta, 2018, 283, 183–189.
crossref
[52] H. Takezawa, S. Ito, H. Yoshizawa and T. Abe, Electrochim. Acta, 2017, 245, 1005–1009.
crossref
[53] M. A. Al-Maghrabi, S. Junji, R. J. Sanderson, V. L. Chevrier, R. A. Dunlap and J. R. Dahn, J. Electrochem. Soc., 2013, 160, A1587–A1593.
crossref
[54] M. K. Kim, B. Y. Jang, J. S. Lee, J. S. Kim and S. Nahm, J. Power Sources, 2013, 244, 115–121.
crossref
[55] H. Takezawa, K. Iwamoto, S. Ito and H. Yoshizawa, J. Power Sources, 2013, 244, 149–157.
crossref
[56] J. Tang, L. Hou, T. Hu, S. Fan, X. Zhou and J. Yang, Compos. Commun., 2021, 23, 100544.
crossref
[57] H.-R. Yang, J. Hwang, H. Seo, K. Kim and J.-H. Kim, J. Power Sources, 2022, 519, 230777.
crossref
[58] J. H. Cho, X. Xiao, M. W. Verbrugge and B. W. Sheldon, ACS Appl. Energy Mater., 2022, 5(11), 13293–13306.
crossref pdf
[59] C. Yidan, R. A. Dunlap and M. N. Obrovac, J. Electrochem. Soc., 2020, 167, 110501.
crossref pdf
[60] Y. Cao, J. C. Bennett, R. A. Dunlap and M. N. Obrovac, Chem. Mater., 2018, 30(21), 7418–7422.
crossref
[61] M. Mamiya, H. Takei, M. Kikuchi and C. Uyeda, J. Cryst. Growth, 2001, 229, 457–461.
crossref
[62] M. Mamiya, M. Kikuchi and H. Takei, J. Cryst. Growth, 2002, 237–239, 1909–1914.

[63] J.-I. Lee, N.-S. Choi and S. Park, Energy Environ. Sci., 2012, 5(7), 7878.
crossref
[64] S. Xu, X. Hou, D. Wang, L. Zuin, J. Zhou, Y. Hou and M. Mann, Adv. Energy Mater., 2022, 12(18), 2200127.

[65] Y. Hwa, C.-M. Park and H.-J. Sohn, J. Power Sources, 2013, 222, 129–134.
crossref
[66] T. Hirose, M. Morishita, H. Yoshitake and T. Sakai, Solid State Ion., 2017, 303, 154–160.
crossref
[67] Y. Nagao, H. Sakaguchi, H. Honda, T. Fukunaga and T. Esaka, J. Electrochem. Soc., 2004, 151(10), A1572.
crossref
[68] M. Miyachi, H. Yamamoto, H. Kawai, T. Ohta and M. Shirakata, J. Electrochem. Soc., 2005, 152(10), A2089.
crossref
[69] T. Morita and N. Takami, J. Electrochem. Soc., 2006, 153(2), A425.
crossref
[70] T. Kim, S. Park and S. M. Oh, J. Electrochem. Soc., 2007, 154(12), A1112.
crossref
[71] J.-H. Kim, C.-M. Park, H. Kim, Y.-J. Kim and H.-J. Sohn, J. Electroanal. Chem., 2011, 661(1), 245–249.
crossref
[72] J. Wang, H. Zhao, J. He, C. Wang and J. Wang, J. Power Sources, 2011, 196(10), 4811–4815.
crossref
[73] M. Yamada, A. Inaba, A. Ueda, K. Matsumoto, T. Iwasaki and T. Ohzuku, J. Electrochem. Soc., 2012, 159(10), A1630–A1635.
crossref
[74] J. H. Yom, S. W. Hwang, S. M. Cho and W. Y. Yoon, J. Power Sources, 2016, 311, 159–166.
crossref
[75] K. Yasuda, Y. Kashitani, S. Kizaki, K. Takeshita, T. Fujita and S. Shimosaki, J. Power Sources, 2016, 329, 462–472.
crossref
[76] T. Huang, Y. Yang, K. Pu, J. Zhang, M. Gao, H. Pan and Y. Liu, RSC Adv., 2017, 7, 2273–2280.
crossref
[77] A. I. Freytag, A. D. Pauric, M. Jiang and G. R. Goward, J. Phys. Chem. C, 2019, 123(18), 11362–11368.
crossref
[78] C.-H. Hsu, H.-Y. Chen and C.-J. Tsai, J. Power Sources, 2019, 438, 226943.
crossref
[79] K. Kitada, O. Pecher, P. C. M. M. Magusin, M. F. Groh, R. S. Weatherup and C. P. Grey, J. Am. Chem. Soc., 2019, 141(17), 7014–7027.
crossref
[80] J. E. Coyle, M. T. Brumbach, G. M. Veith and C. A. Apblett, J. Phys. Chem. C, 2020, 124(15), 8153–8161.
crossref
[81] K. W. Kim, H. Park, J. G. Lee, J. Kim, Y.-U. Kim, J. H. Ryu, J. J. Kim and S. M. Oh, Electrochim. Acta, 2013, 103, 226–230.
crossref
[82] Z. Xiao, C. Yu, X. Lin, X. Chen, C. Zhang and F. Wei, Carbon, 2019, 149, 462–470.
crossref
[83] H. Li and H. Zhou, Chem. Commun., 2012, 48(9), 1201–1217.
crossref
[84] F. Wang, G. Chen, N. Zhang, X. Liu and R. Ma, Carbon Energy, 2019, 1(2), 219–245.
crossref pdf
[85] C.-H. Doh, H.-M. Shin, D.-H. Kim, Y.-C. Ha, B.-S. Jin, H.-S. Kim, S.-I. Moon and A. Veluchamy, Electrochem. Commun., 2008, 10(2), 233–237.
crossref
[86] W.-R. Liu, Y.-C. Yen, H.-C. Wu, M. Winter and N.-L. Wu, J. Appl. Electrochem., 2009, 39, 1643–1649.
crossref pdf
[87] M. Yoshio, H. Wang, K. Fukuda, T. Umeno, N. Dimov and Z. Ogumi, J. Electrochem. Soc., 2002, 149(12), A1598.
crossref
[88] S. Jeong, X. Li, J. Zheng, P. Yan, R. Cao, H. J. Jung, C. Wang, J. Liu and J.-G. Zhang, J. Power Sources, 2016, 329, 323–329.
crossref
[89] S. J. Lee, H. J. Kim, T. H. Hwang, S. Choi, S. H. Park, E. Deniz, D. S. Jung and J. W. Choi, Nano Lett., 2017, 17(3), 1870–1876.
crossref
[90] Z. Li, Q. He, L. He, P. Hu, W. Li, H. Yan, X. Peng, C. Huang and L. Mai, J. Mater. Chem. A, 2017, 5(8), 4183–4189.
crossref
[91] L. Zhai, C. Hou, H. Sun, Y. Qu, J. Hou, H. Lu, J. Wu, B. Wang, Y. Ma and X. Guo, J. Energy Storage, 2024, 97(Part A), 112715.
crossref
[92] Z. You, C. Lin, P. Zheng, J. Li, Q. Feng, J. Tao, Y. Zheng, Y. Lin, Z. Huang and J. Li, Small, 2025, 21(2), 2407215.

[93] H. Xie, C. Hou, Z. Yue, L. Zhai, H. Sun, H. Lu, J. Wu, S. Yang and Y. Ma, J. Energy Storage, 2023, 64, 107147.
crossref
[94] S. Yi, M. Su, Z. Yan, Y. Zhang, Y. Xiao, X. Zhang, Z. Wang, P. Ning, D. Yang and N. Du, Carbon, 2024, 228, 119367.
crossref
[95] X. Zhang, M. Zhao, J. Zhang and W. Wang, J. Power Sources, 2025, 625, 235680.
crossref
[96] X. Zhang, M. Zhao, C. M. Costa, J. Zhang, F. Scheiba, S. Lanceros-Méndez, W. Wang and Q. Zhang, Small Methods, 2025, 9(7), 2401829.

[97] Z. Liu, Y. Zhao, R. He, W. Luo, J. Meng, Q. Yu, D. Zhao, L. Zhou and L. Mai, Energy Storage Mater., 2019, 19, 299–305.
crossref
[98] M. Li, Y. Zeng, Y. Ren, C. Zeng, J. Gu, X. Feng and H. He, J. Power Sources, 2015, 288, 53–61.
crossref
[99] Z. Bao, M. R. Weatherspoon, S. Shian, Y. Cai, P. D. Graham, S. M. Allan, G. Ahmad, M. B. Dickerson, B. C. Church, Z. Kang, C. J. Summers, M. Liu and K. H. Sandhage, Nature, 2007, 446, 172–175.
crossref pdf
[100] M.-S. Wang, L.-Z. Fan, M. Huang, J. Li and X. Qu, J. Power Sources, 2012, 219, 29–35.
crossref
[101] D. S. Jung, M.-H. Ryou, Y. J. Sung, S. B. Park and J. W. Choi, Proc. Natl. Acad. Sci. U.S.A., 2013, 110, 12229–12234.
crossref
[102] J. Cui, Y. Cui, S. Li, H. Sun, Z. Wen and J. Sun, ACS Appl. Mater. Interfaces, 2016, 8(44), 30239–30247.
crossref
[103] Y. Ju, J. A. Tang, K. Zhu, Y. Meng, C. Wang, G. Chen, Y. Wei and Y. Gao, Electrochim. Acta, 2016, 191, 411–416.
crossref
[104] G. Zhu, F. Zhang, X. Li, W. Luo, L. Li, H. Zhang, L. Wang, Y. Wang, W. Jiang, H. K. Liu, S. X. Dou and J. Yang, Angew. Chem. Int. Ed., 2019, 58(20), 6669–6673.
crossref pdf
[105] Y. Wang, C. Xu, J. Fu, S. Che, G. Ma, K. Chen, J. Yang, S. Liu, G. Cheng, D. Zhang, Y. Huang, W. Gao, Y. Cui and Y. Li, Chem. Eng. J., 2025, 510, 161895.
crossref
[106] Y. Xiao, Y. Mao, T. Li, X. Hao and W. Wang, ACS Appl. Mater. Interfaces, 2023, 15(39), 45938–45948.
crossref pdf
[107] L. Lee, W. To A Ran, J.-H. Lee, S. M. Hwang and Y.-J. Kim, Chem. Eng. J, 2022, 442(Part 1), 136166.
crossref
[108] Z. Zhou, Z. Li, X. Liu, J. Wu, K. Li and C. Wang, Chem. Eng. J., 2025, 511, 162297.
crossref
[109] M. Li, J. Qiu, H. Ming, P. Zhao, Z. Jin, S. Zhang and Y. Yang, J. Alloys Compd., 2019, 809, 151659.
crossref
[110] Z. Gu, W. Li, Y. Chen, X. Xia and H. Liu, Sustain. Energy Fuels, 2020, 4(5), 2333–2341.
crossref
[111] B. Liu, A. Abouimrane, D. E. Brown, X. Zhang, Y. Ren, Z. Z. Fang and K. Amine, J. Mater. Chem. A, 2013, 1(13), 4376–4382.
crossref
[112] J. Zhang, J. Zhang, T. Bao, X. Xie and B. Xia, J. Power Sources, 2017, 348, 16–20.
crossref
[113] Z. Li, H. Zhao, P. Lv, Z. Zhang, Y. Zhang, Z. Du, Y. Teng, L. Zhao and Z. Zhu, Adv. Funct. Mater., 2018, 28(31), 1605711.

[114] H. Xue, Y. Wu, Y. Zou, Y. Shen, G. Liu, Q. Li, D. Yin, L. Wang and J. Ming, Adv. Funct. Mater., 2020, 30(21), 1910657.

[115] Y. Jiang, S. Liu, Y. Ding, J. Jiang, W. Li, S. Huang, Z. Chen, B. Zhao and J. Zhang, J. Power Sources, 2020, 467, 228301.
crossref
[116] G. Jeong, J.-H. Kim, Y.-U. Kim and Y.-J. Kim, J. Mater. Chem., 2012, 22(16), 7999–8004.
crossref
[117] J. Bae, D. S. Kim, H. Yoo, E. Park, Y.-G. Lim, M.-S. Park, Y.-J. Kim and H. Kim, ACS Appl. Mater. Interfaces, 2016, 8(7), 4541–4547.
crossref
[118] P. Wang, S. Hou, F. Pang, M. Liu, Y. Li, T. Liu, Y. Luo, Y. Fan and L. Zhao, ChemElectr°Chem, 2019, 6(5), 1574–1581.
crossref pdf
[119] G. Jeong, Y.-U. Kim, S. A. Krachkovskiy and C. K. Lee, Chem. Mater., 2010, 22(19), 5570–5579.
crossref
[120] Y. Zhang, G. Guo, C. Chen, Y. Jiao, T. Li, X. Chen, Y. Yang, D. Yang and A. Dong, J. Power Sources, 2019, 426, 116–123.
crossref
[121] W. He, Y. Liang, H. Tian, S. Zhang, Z. Meng and W.-Q. Han, Energy Storage Mater., 2017, 8, 119–126.
crossref
[122] K. Kim, H. Choi and J.-H. Kim, Appl. Surf. Sci., 2017, 416, 527–535.
crossref
[123] R. Fu, Y. Wu, C. Fan, Z. Long, G. Shao and Z. Liu, ChemSusChem, 2019, 12(14), 3377–3382.
crossref pdf
[124] J. Zhao, L. Shi, Z. Lang, G. Jia, D. Lan, Y. Cui and J. Cui, J. Electroanal. Chem., 2025, 977, 118878.
crossref
[125] Y. Ling, T. Chen, S. Chen, B. Wang, P. Zeng, S. Shen, C. Yuan, Z. Zhou, J. Wang and L. Zhang, ACS Appl. Energy Mater., 2025, 8(10), 6723–6732.
crossref pdf
[126] D. Youn, N. G. Kim, W. J. Jeong, D. J. Chung, J. Y. Kim and H. Kim, ACS Appl. Mater. Interfaces, 2022, 14(40), 45333–45341.
crossref pdf
[127] D. Youn, N. G. Kim, J. Y. Kim, S. T. Kim, D. J. Chung and H. Kim, J. Energy Storage, 2024, 89, 111641.
crossref
[128] D. S. Kim, Y. E. Kim and H. Kim, J. Power Sources, 2019, 422, 18–24.
crossref
[129] J.-M. Lee, H. Jung, Y. Hwa, H. Kim, D. Im, S.-G. Doo and H.-J. Sohn, J. Power Sources, 2010, 195(15), 5044–5048.
crossref
[130] F. Holtstiege, P. Bärmann, R. Nölle, M. Winter and T. Placke, Batteries, 2018, 4(1), 4.
crossref
[131] H. J. Kim, S. Choi, S. J. Lee, M. W. Seo, J. G. Lee, E. Deniz, Y. J. Lee, E. K. Kim and J. W. Choi, Nano Lett., 2016, 16(1), 282–288.
crossref
[132] H. Zhao, Z. Wang, P. Lu, M. Jiang, F. Shi, X. Song, Z. Zheng, X. Zhou, Y. Fu, G. Abdelbast, X. Xiao, Z. Liu, V. S. Battaglia, K. Zaghib and G. Liu, Nano Lett., 2014, 14(11), 6704–6710.
crossref
[133] E. Park, D. J. Chung, M.-S. Park and H. Kim, J. Power Sources, 2019, 440, 227094.
crossref
[134] Q. Meng, G. Li, J. Yue, Q. Xu, Y.-X. Yin and Y.-G. Guo, ACS Appl. Mater. Interfaces, 2019, 11(35), 32062–32068.
crossref
[135] T. Tabuchi, H. Yasuda and M. Yamachi, J. Power Sources, 2005, 146(1–2), 507–509.
crossref
[136] M.-Y. Yan, G. Li, J. Zhang, Y.-F. Tian, Y.-X. Yin, C.-J. Zhang, K.-C. Jiang, Q. Xu, H.-L. Li and Y.-G. Guo, ACS Appl. Mater. Interfaces, 2020, 12, 27202–27209.
crossref
[137] I. W. Seong, K. T. Kim and W. Y. Yoon, J. Power Sources, 2009, 189, 511–514.
crossref
[138] B. Huang, T. Huang, L. Wan and A. Yu, ACS Sustain. Chem. Eng., 2021, 9(2), 648–657.
crossref
[139] X. Li, C. Bian, J. Zhang, J. Hong, R. Fu, X. Zhou, Z. Liu and G. Shao, ACS Appl. Energy Mater., 2023, 6, 8919–8928.
crossref pdf
[140] Y. Li, Y. Qian, B. Deng and N. Lin, Chem. Eng. J., 2024, 488, 151012.
crossref
[141] D. I. Lee, H.-W. Yang, W. S. Kang, J. Kim and S.-J. Kim, J. Electrochem. Soc., 2019, 166, A787.
crossref
[142] H. Takezawa, S. Ito, H. Yoshizawa and T. Abe, Chem. Lett., 2017, 46(9), 1365–1367.
crossref
[143] A. Veluchamy, C.-H. Doh, D.-H. Kim, J.-H. Lee, D.-J. Lee, K.-H. Ha, H.-M. Shin, B.-S. Jin, H.-S. Kim, S.-I. Moon and C.-W. Park, J. Power Sources, 2009, 188, 574–577.
crossref
[144] J. Zhao, H.-W. Lee, J. Sun, K. Yan, Y. Liu, W. Liu, Z. Lu, D. Lin, G. Zhou and Y. Cui, Proc. Natl. Acad. Sci. U.S.A., 2016, 113, 7408–7413.
crossref
[145] X. Yang, Z. Wen, X. Xu, B. Lin and S. Huang, J. Power Sources, 2007, 164(2), 880–884.
crossref
[146] D. J. Chung, D. Youn, S. Kim, D. Ma, J. Lee, W. J. Jeong, E. Park, J.-S. Kim, C. Moon, J. Y. Lee, H. Sun and H. Kim, Nano Energy, 2021, 89(Part A), 106378.
crossref
[147] D. J. Chung, D. Youn, J. Y. Kim, W. J. Jeong, S. Kim, D. Ma, T. R. Lee, S. T. Kim and H. Kim, Small, 2022, 18(27), 2202209.

[148] C. Gao, H. Zhao, P. Lv, C. Wang, J. Wang, T. Zhang and Q. Xia, J. Electrochem. Soc., 2014, 161, A2216–A2221.
crossref
[149] B.-C. Yu, Y. Hwa, J.-H. Kim and H.-J. Sohn, Electrochim. Acta, 2014, 117, 426–430.
crossref
[150] C.-K. Back, T.-J. Kim and N.-S. Choi, J. Mater. Chem. A, 2014, 2(33), 13648.
crossref
[151] J.-I. Lee, K. T. Lee, J. Cho, J. Kim, N.-S. Choi and S. Park, Angew. Chem. Int. Ed. Engl., 2012, 51(11), 2767–2771.
crossref
[152] J.-I. Lee and S. Park, Nano Energy, 2013, 2(1), 146–152.
crossref
[153] D. Park, H.-S. Kim, H. Seo, K. Kim and J.-H. Kim, Electrochim. Acta, 2020, 357, 136862.
crossref
[154] A. Raza, J. Y. Jung, C.-H. Lee, B. G. Kim, J.-H. Choi, M.-S. Park and S.-M. Lee, ACS Appl. Mater. Interfaces, 2021, 13(6), 7161–7170.
crossref
[155] E. Park, H. Yoo, J. Lee, M.-S. Park, Y.-J. Kim and H. Kim, ACS Nano, 2015, 9, 7690–7696.
crossref
[156] T. Zhang, J. Gao, H. P. Zhang, L. C. Yang, Y. P. Wu and H. Q. Wu, Electrochem. Commun., 2007, 9, 886–890.
crossref
[157] F. Dai, R. Yi, M. L. Gordin, S. Chen and D. Wang, RSC Adv., 2012, 2, 12710–12713.
crossref
[158] C. K. Chan, H. Peng, G. Liu, K. McIlwrath, X. F. Zhang, R. A. Huggins and Y. Cui, Nat. Nanotechnol., 2008, 3, 31–35.
crossref pdf
[159] T. Song, J. Xia, J.-H. Lee, D. H. Lee, M.-S. Kwon, J.-M. Choi, J. Wu, S. K. Doo, H. Chang, W. I. Park, D. S. Zang, H. Kim, Y. Huang, K.-C. Hwang, J. A. Rogers and U. Paik, Nano Lett., 2010, 10, 1710–1716.
crossref
[160] H. Guo, R. Mao, X. Yang and J. Chen, Electrochim. Acta, 2012, 74, 271–274.
crossref
[161] H. Yoo, J.-I. Lee, H. Kim, J.-P. Lee, J. Cho and S. Park, Nano Lett., 2011, 11(10), 4324–4328.
crossref
[162] S. Yoo, J.-I. Lee, M. Shin and S. Park, ChemSusChem, 2013, 6(7), 1153–1157.
crossref
[163] J. Lee, J. Moon, S. A. Han, J. Kim, V. Malgras, Y.-U. Heo, H. Kim, S.-M. Lee, H. K. Liu, S. X. Dou, Y. Yamauchi, M.-S. Park and J. H. Kim, ACS Nano, 2019, 13(8), 9607–9619.
crossref
[164] J. Yu, C. Zhang, X. Huang, L. Cao, A. Wang, W. Dai, D. Li, Y. Dai, C. Zhou, Y. Zhang and Y. Zhang, Nanomaterials, 2024, 14(14), 1223.
crossref


ABOUT
ARTICLE CATEGORY

Browse all articles >

BROWSE ARTICLES
AUTHOR INFORMATION
Editorial Office
E-mail: journal@kecs.or.kr    Tel: +82-2-568-9392    Fax: +82-2-568-5931                   

Copyright © 2025 by The Korean Electrochemical Society.

Developed in M2PI

Close layer
prev next