Next Article in Journal
State of Health Estimation of Lithium-Ion Battery Based on Novel Health Indicators and Improved Support Vector Regression
Previous Article in Journal
From Thermosetting Resins to Energy Devices: A Review on Polybenzoxazine-Derived Materials for Supercapacitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Various Technologies to Mitigate Volume Expansion of Silicon Anode Materials in Lithium-Ion Batteries

Department of Energy Convergence Engineering, Cheongju University (CJU), 111 Taejeong-ro, Maengdong-myeon, Eumseong-gun 27739, Chungcheongbuk-do, Republic of Korea
*
Author to whom correspondence should be addressed.
Batteries 2025, 11(9), 346; https://doi.org/10.3390/batteries11090346
Submission received: 22 August 2025 / Revised: 10 September 2025 / Accepted: 12 September 2025 / Published: 21 September 2025

Abstract

Silicon anodes for lithium-ion batteries (LIBs) offer exceptional theoretical capacity (~4200 mAh/g) but face critical challenges due to significant volume expansion (>300%) during lithiation, leading to mechanical degradation and rapid capacity fading. This review highlights recent advancements in mitigating these issues, including structural designs such as core–shell architectures, porous composites, and multidimensional encapsulation techniques that buffer mechanical stress and stabilize the solid electrolyte interphase (SEI). Binder innovations and hybrid material systems further enhance electrode integrity and cycling stability. While substantial progress has been made, challenges remain in scaling these solutions for commercial applications. This paper provides insights into current strategies and future directions for enabling silicon-based anodes in next-generation LIBs.

Graphical Abstract

1. Potential and Challenges of Silicon Anodes

While lithium-ion batteries (LIBs) play a central role in modern society’s energy transition, currently commercialized graphite anodes have reached their theoretical capacity limit (372 mAh/g), creating a major obstacle for further energy density enhancement [1]. To overcome this limitation, researchers are actively investigating high-capacity anode active materials such as silicon (Si), tin (Sn), and metal oxides [2,3]. Silicon, in particular, offers a theoretical capacity more than 10 times higher than graphite, but faces challenges related to volume expansion during charge–discharge cycles and the resulting electrode structural instability. Therefore, multifaceted approaches including nanostructuring, hybridization, and surface modification are being pursued to maximize the potential of silicon anode materials [4,5], which is gaining attention as a key research direction for building sustainable future energy systems.
Beyond these large-scale applications, silicon anode technology is also paving the way for the development of compact and lightweight three-dimensional (3D) lithium-ion microbatteries for on-chip power supplies. In this emerging field, aligned silicon vertical pillar arrays, fabricated directly on silicon substrates using mature semiconductor processes, enable the monolithic integration of battery anodes with electronic circuits and MEMS. Interestingly, even in this micro-scale domain, hybrid material system approaches, such as applying a carbon coating to silicon nanowires to suppress volume expansion, are being actively investigated as a key strategy. For a comprehensive overview of this specific area, readers are referred to recent reviews on micro lithium batteries, advances in 3D silicon-based microbatteries, and the effects of carbon-coating on structured Si nanowires [6,7,8].
The aforementioned endeavors to replace conventional graphite anodes have placed Si in the spotlight as the most promising next-generation anode material [9]. Si boasts a theoretical mass-specific capacity of approximately 4200 mAh/g (based on Li4.4Si), which is more than 10 times higher than that of graphite (372 mAh/g), a relatively low operating potential (approximately 0.4 V vs. Li/Li+), along with abundant natural reserves and environmental friendliness [10,11]. Figure 1 schematically illustrates the morphological evolution of silicon microparticles and nanoparticles during repeated lithiation and delithiation. Silicon undergoes extreme volumetric expansion of up to ~400% during the formation of LixSi alloys, which generates significant internal stress. This stress leads to particle cracking, pulverization, and continuous rupture of the solid electrolyte interphase (SEI), accelerating capacity fading and impedance growth. The figure also highlights the difference between micron-sized silicon, which tends to fail catastrophically, and nanosized silicon, which better accommodates volume changes due to shorter diffusion paths and higher surface-to-volume ratios. These insights emphasize why nanostructuring and composite architectures are crucial to enhance mechanical stability and cycling durability.
Reinforcing its sustainability, recent studies have demonstrated viable methods for recovering high-performance, porous micro-sized silicon from waste AlSi alloys, highlighting a promising direction for cost-effective and eco-friendly production [12]. The successful commercialization of silicon anodes could dramatically increase the energy density of LIBs, enabling significantly extended driving range for electric vehicles and prolonged usage time for portable devices. However, silicon anodes face serious technical challenges, primarily the extreme volume change during alloying and dealloying reactions with lithium ions [13,14]. Silicon can expand by up to 400% when forming LixSi alloys, causing substantial mechanical stress that leads to cracking and pulverization of active material particles [15,16], weakened adhesion between components resulting in delamination from the current collector [17], and formation of thick, unstable solid electrolyte interphase (SEI) layers at newly exposed surfaces that impede ion movement and cause capacity loss [18]. These complex problems severely degrade cycle life and electrical conductivity, creating major obstacles for commercialization [19]. Therefore, innovative solutions to control volume expansion and mitigate secondary issues are urgently needed to utilize silicon anodes’ inherent advantages.
Addressing the low initial Coulombic efficiency (ICE) is also a critical task, and recent interfacial engineering on porous SiOx@C composites has demonstrated the possibility of achieving high ICE (93.4%) through prelithiation, offering a promising solution to this long-standing problem [20].
Figure 1. Schematic illustrations of morphology variation in Si active materials caused by charge/discharge processes. The Si active materials were (a) Si microparticles and (b) Si nanoparticle aggregates. Reproduced from ref. [15], licensed under CC BY 4.0 (https://creativecommons.org/licenses/by/4.0/). Accessed on 30 May 2025.
Figure 1. Schematic illustrations of morphology variation in Si active materials caused by charge/discharge processes. The Si active materials were (a) Si microparticles and (b) Si nanoparticle aggregates. Reproduced from ref. [15], licensed under CC BY 4.0 (https://creativecommons.org/licenses/by/4.0/). Accessed on 30 May 2025.
Batteries 11 00346 g001

2. Concept and Principles of Hybrid Material Systems

2.1. Definition of Hybrid Material Systems

Hybrid material systems combine two or more different materials at the nanometer or micrometer level to achieve enhanced properties that single components cannot provide alone [21,22]. These systems maximize the advantages of each constituent material while compensating for individual disadvantages through synergistic effects [23].
In silicon anodes, hybrid material systems offer critical solutions to address volume expansion issues [24]. Figure 2 details a multi-step mechanochemical process starting from micrometric Si and CuO. Through high-energy and wet milling, these precursors are transformed into a composite of nanocrystalline and amorphous phases. A final carbonization step with glucose yields the carbon-coated Si/Cu3Si composite. This illustrates a practical approach to creating a hybrid material designed to buffer volume changes and improve conductivity simultaneously. While silicon offers high theoretical capacity, its extreme volume expansion causes cracking, delamination, unstable SEI formation, and capacity/conductivity loss [16]. Hybrid material systems provide multifaceted solutions to mitigate these problems while improving electrical conductivity and interfacial stability. Key examples include Si/C composites that provide mechanical buffering and enhanced conductivity, Si/graphene systems that offer excellent electrical pathways and structural support, and Si/polymer binder combinations that maintain electrode integrity during cycling. These approaches represent key technological strategies for silicon anode commercialization [25].

2.2. Working Mechanisms of Hybrid Material Systems in Silicon Anodes

Hybrid material systems address silicon anode challenges through four core mechanisms that work synergistically to overcome the fundamental limitations of pure silicon [26,27]. The most critical function involves volume expansion mitigation, where core–shell structures with silicon nanoparticles encapsulated by carbon shells physically constrain expansion [28], while yolk–shell structures with intentional void spaces allow silicon expansion without electrode deformation [29]. Figure 3 illustrates how yolk–shell design preserves electrode integrity, showing that at low current density (10mC), the crystalline Si core experiences significant stress leading to cracking, but at higher density (50mC), the Si core becomes amorphous while the overall structure remains intact due to adequate void space for volume accommodation.
Equally important is the enhancement of electrical conductivity throughout the electrode structure. Carbon-based materials such as graphene, CNTs, and amorphous carbon serve as conductive matrices providing electron pathways between silicon particles and current collectors [30,31]. Conductive polymers and certain metal oxides also contribute to conductivity improvement [32], creating a robust electrical network that maintains connectivity even during the dramatic volume changes experienced by silicon particles.
The stabilization of the solid electrolyte interphase represents another crucial mechanism for improving silicon anode performance. Protective coating layers including carbon, Al2O3, and TiO2 prevent direct silicon–electrolyte contact, stabilizing SEI formation and reducing electrolyte consumption [29,33,34]. This approach improves ICE and cycle stability [35,36], while advanced approaches include fluorinated MXene coatings that create LiF-rich, stable SEI layers [37]. These protective layers act as selective barriers that allow lithium ion transport while preventing unwanted side reactions.
Finally, mechanical stability enhancement plays a vital role in maintaining electrode integrity over extended cycling. Carbon matrices and metal oxide shells provide superior mechanical strength and flexibility compared to silicon, preventing particle pulverization and maintaining electrode integrity through strong interfacial interactions [19,30,38]. This mechanical reinforcement works in concert with the other mechanisms to create a robust electrode structure that can withstand the stresses associated with repeated lithiation and delithiation cycles.

3. Types and Characteristics of Hybrid Material Systems

Hybrid material systems can be classified in various ways according to their constituent materials and structural characteristics. The hybrid systems that have been most actively researched for improving the performance of silicon anodes can be broadly categorized by their constituent materials into silicon–carbon, silicon–metal oxide, silicon–2D material, and silicon–conductive polymer hybrid systems, among others. This section examines in detail the structure, characteristics, performance, and latest research trends of each of these types of hybrid systems. Furthermore, synergistic strategies combining different modification approaches, such as germanium (Ge) doping to enhance conductivity and liquid metal alloy surface modification to create a flexible buffer layer, are being explored to maximize the performance of micron-sized silicon anodes [39].

3.1. Classification According to Constituent Materials

3.1.1. Silicon–Carbon Hybrid Systems

Carbon-based materials are the most widely researched hybrid partners for improving the performance of silicon anodes due to their excellent electrical conductivity, high mechanical stability, chemical stability, and the various possibilities they offer for structural control [40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69]. Hybrid systems that are a combination of silicon and carbon offer multifaceted benefits: the carbon matrix effectively buffers the volume expansion of silicon, provides excellent electrical conduction pathways, and facilitates the formation of stable SEI layers. Representative silicon–carbon hybrid systems include composites with graphene [52], CNTs [53], and carbon nanofibers (CNFs) [54]. Figure 4 provides a comprehensive overview of the major structural designs for Si/C anodes. It categorizes the approaches into Embedded Structures, where Si is dispersed in a carbon matrix; Carbon-coated Structures, featuring a core–shell design; and Hollow Structures, which incorporate void spaces. The accompanying TEM images for each category offer a visual confirmation of these nanoscale architectures, showcasing the diverse strategies employed to tackle the volume expansion issue.
Graphene, a two-dimensional carbon material with outstanding electrical conductivity, a large specific surface area, excellent mechanical strength, and flexibility, has attracted attention as an ideal material for hybridization with silicon [55]. Various structures have been studied, including silicon nanoparticles wrapped with graphene sheets, sandwich structures in which silicon particles are intercalated between graphene sheets, or silicon dispersed within three-dimensional graphene networks [56]. Graphene effectively accommodates the volume expansion of silicon, prevents particle agglomeration, and provides rapid electron transfer pathways, which, in combination, result in high capacity and excellent cycle and rate characteristics [57]. For example, hybrid electrodes fabricated by depositing silicon onto three-dimensional porous graphene foam using chemical vapor deposition (CVD) demonstrated high reversible capacity and long-term cycle stability [58].
Carbon nanotubes (CNTs), as one-dimensional carbon materials with high aspect ratios, excellent electrical conductivity, and mechanical strength, have also been actively applied to silicon hybrids. CNTs form entangled networks around silicon particles to provide effective electron conduction pathways and disperse the mechanical stress that accumulates during the volume changes experienced by Si. Additionally, CNT networks contribute to enhancing the structural stability of electrodes by strengthening the adhesion between active materials, conductive additives, and binders within the electrode [59]. CNT-Si hybrid materials are produced using methods such as directly growing silicon nanoparticles on CNT surfaces or mixing silicon and CNTs to manufacture composites, and hybrid electrodes utilizing CNTs have been reported to exhibit improved rate capabilities and cycle stability [60].
Carbon nanofibers (CNFs), manufactured using methods such as electrospinning, have also been usefully employed in silicon hybrid systems. CNFs have continuous one-dimensional structures and large specific surface areas and can be hybridized by either incorporating silicon nanoparticles within their porous nanofibers or by coating their surfaces with silicon. CNFs effectively buffer the volume expansion of Si through their flexible structure, and the fiber network provides excellent pathways for electron and ion transfer [61]. In particular, structures comprising silicon embedded inside porous CNFs accommodate the volume expansion within the internal space to minimize external deformation and induce stable SEI formation, thus improving the electrochemical performance [62].
Researchers working on these various silicon–carbon hybrid systems have reported significant performance improvements in the initial capacity, Coulombic efficiency (CE), cycle stability, and rate characteristics compared to pure silicon or conventional graphite anodes [63]. A prime example of such improvements is detailed in Figure 5. The schematic in Figure 5a illustrates the synthesis of crumpled graphene-encapsulated silicon nanoparticles (cpDOPA-crGO-Si) via a spray-drying process. As quantitatively demonstrated in Figure 5b, this graphene-encapsulated structure maintained a high capacity of over 1000 mAh/g even after 100 cycles at a current density of 1 A/g [63]. This result validates that the graphene encapsulation strategy effectively buffers volume changes and ensures stable, long-term performance. Meanwhile, other approaches, such as hierarchically combined porous silicon structures with CNTs, have also exhibited excellent capacity retention rates even at high rates of 5 A/g [64]. Each hybrid system has unique advantages and disadvantages depending on the type of carbon material, structure, and bonding method with silicon, and requires the system design to be optimized according to the performance objectives required for specific applications.
Looking at recent research trends, ternary carbon-based hybrid systems, which maximize the synergistic effects beyond the performance of single carbon materials by using two or more types of carbon materials together, are being actively studied. For example, the purpose of structures that simultaneously use graphene and CNTs to wrap silicon particles combine the two-dimensional coverage provided by graphene with the one-dimensional connectivity of the CNTs to more effectively buffer the volume expansion and construct the electron transfer network [65]. Additionally, multifunctional carbon composites, which offer enhanced electrical conductivity, surface affinity, and lithium storage capability by accepting heteroatoms such as nitrogen (N), sulfur (S), and phosphorus (P) as dopants into their carbon matrices, are also attracting attention [66,67]. These latest developments are contributing to further enhancing the performance of silicon–carbon hybrid systems and increasing their potential for commercialization. Table 1 summarizes the electrochemical performance of various silicon-carbon hybrid systems, presenting key performance metrics including current density, cycle number, and remaining capacity. This data quantitatively demonstrates the effectiveness of carbon-based hybrid systems in enhancing silicon anode performance.
Despite the numerous advantages, silicon–carbon hybrid systems face several challenges that hinder their widespread commercialization. The high cost and complex manufacturing processes of high-quality carbon materials like graphene and CNTs are significant economic barriers. Achieving a uniform and scalable carbon coating on silicon nanoparticles is technically demanding, and batch-to-batch consistency remains a major hurdle for industrial production [68,69].
Furthermore, there is an inherent trade-off between the carbon content and the overall energy density of the electrode. While a sufficient amount of carbon is necessary to ensure good electrical conductivity and buffer the volume expansion, an excessive carbon content reduces the proportion of active silicon, thereby lowering the specific capacity. The long-term stability of the Si-C interface during repeated cycling is another concern, as mechanical stress can lead to delamination and loss of electrical contact [38]. Finally, the potential for carbon oxidation at high operating voltages and the environmental impact associated with the production and disposal of carbon nanomaterials need to be carefully considered.

3.1.2. Silicon–Metal Oxide Hybrid Systems

In addition to carbon-based materials, various metal oxides (MOx) are being actively researched as hybrid partners to improve the performance of silicon anodes [70]. MOx exhibit semiconducting, insulating, or ionic conductive properties depending on their chemical composition and some possess inherent lithium storage capabilities [71]. The hybridization of Si with MOx is beneficial in various ways: metal oxide layers can act as mechanical protective membranes to buffer the volume expansion of Si, provide surface modification effects that stabilize SEI formation, and in some cases, offer additional capacity. Notably, TiO2 [34], Fe2O3 [70], and SnO2 [72] have been incorporated into hybrid systems with silicon.
Titanium dioxide (TiO2) is known for its chemical stability, excellent mechanical strength, and relatively good lithium ion conductivity [73]. Hybrid structures that have been studied include coating silicon nanoparticles with thin layers of TiO2 [74] or inserting silicon into TiO2 nanostructures. TiO2 coating layers prevent direct contact between the silicon surface and electrolyte, which inhibits side reactions and induces stable SEI formation [75]. They also serve as mechanical buffer layers by contributing to mitigating the structural destruction caused by the volume expansion of Si [76]. For example, hybrid electrodes coated with uniform TiO2 layers on silicon nanowire surfaces using atomic layer deposition (ALD) techniques demonstrated improved cycle stability and CE compared to uncoated silicon nanowires [77].
Iron oxide (Fe2O3) is a conversion reaction-based anode active material with a high theoretical capacity (approximately 1007 mAh/g), which can contribute to increasing the overall electrode capacity when composited with silicon [78]. Additionally, Fe2O3 has excellent mechanical strength, which, as would be anticipated, suppress the volume expansion of Si. Various approaches to preparing Fe2O3–Si composites have been studied, including mixing silicon nanoparticles with Fe2O3 nanoparticles [70], applying a coating of Fe2O3 to silicon surfaces [79], or dispersing silicon within Fe2O3 matrices [80]. However, Fe2O3 itself undergoes substantial volume changes during charging and discharging and has low electrical conductivity, which has resulted in parallel research to secure electrical conductivity and structural stability by additionally introducing carbon coatings and other measures [81]. Figure 6 provides a clear example of this ternary hybrid strategy. The synthesis schematic in Figure 6a shows Si nanoparticles being coated first with a metal–organic framework (MIL-88-Fe) template, which is then annealed to form a dual shell of Fe2O3 and carbon. The cycling performance graph in Figure 6b powerfully illustrates the synergy of this structure. While both pristine Si and Si@C anodes show rapid capacity decay, the Si@Fe2O3/C composite anode delivers a stable capacity of nearly 1000 mAh/g over 400 cycles. This demonstrates that the dual-shell approach, combining a mechanically robust metal oxide with a conductive carbon layer, is highly effective for achieving long-term stability.
Tin dioxide (SnO2) has also been used as an anode active material with a high theoretical capacity (approximately 782 mAh/g, based on alloying reactions), which can increase the energy density of electrodes through hybridization with silicon. The electrical conductivity of SnO2 is also superior to that of silicon, thus improved electron transfer effects can be expected when composited [82]. The formation of SnO2 shells over silicon cores or fabrication of silicon composites with SnO2 nanostructures has also been studied [83,84,85]. However, since SnO2 also undergoes considerable volume changes during charging and discharging, it is important to secure structural stability through nanostructuring or carbon compositing [86].
Si-MOx hybrid systems exhibit various performance improvement effects depending on the type and structure of the metal oxide. While TiO2 coatings primarily focus on enhancing the interfacial stability, Fe2O3 or SnO2 composites could be expected to simultaneously offer capacity enhancement effects. However, the low electrical conductivity and problematic volume change in the MOx themselves are challenges that remain to be solved. Recent research has therefore focused on maximizing the advantages of each material by coating multiple layers of two or more types of MOx to extend the benefits beyond those of a single metal oxide [71], or developing smart composites that can maintain the structural stability despite repetitive volume changes by imparting self-healing functions to MOx [87]. Additionally, research on ternary hybrid systems utilizing both MOx and carbon materials is being actively conducted, with efforts being made to comprehensively improve the electrical conductivity, mechanical stability, and interfacial stability [88].
While metal oxide coatings offer significant benefits, they also present several challenges. Many metal oxides have low intrinsic electrical conductivity, which can increase the overall impedance of the electrode and limit the rate capability. Although they can act as an artificial SEI, some metal oxides may still react with the electrolyte over time, leading to capacity fading [38]. The coating process itself, often involving techniques like atomic layer deposition (ALD) [89] or sol–gel methods, can be expensive and difficult to scale up.
Moreover, the mechanical properties of the metal oxide layer are crucial. If the coating is too rigid, it can fracture during the volume expansion of silicon, exposing the active material to the electrolyte. Conversely, if it is too soft, it may not provide sufficient mechanical constraint. Achieving the optimal thickness, uniformity, and mechanical properties of the metal oxide coating is therefore a key challenge in this hybrid system. Table 2 presents the electrochemical performance data of hybrid systems combining silicon with various metal oxides (Fe2O3, SnO2, TiO2, etc.). The numerical data confirms the contribution of metal oxide coatings to mitigating silicon volume expansion and improving interfacial stability.

3.1.3. Silicon–2D Material Hybrid Systems

Following the successful application of graphene, various two-dimensional (2D) nanomaterials beyond graphene have been receiving significant attention due to their unique physicochemical properties, and their potential as hybrid partners for improving the performance of silicon anodes is being actively explored [90]. Two-dimensional materials exhibit atomic-level thinness, a large specific surface area, excellent mechanical flexibility and strength, and diverse electrical properties (metallic, semiconducting, insulating) depending on their type. These favorable characteristics can effectively buffer the volume expansion of Si, optimize electron and ion transfer pathways, and enhance the interfacial stability. Representative examples that are being applied to hybrid systems with silicon include MXenes [90,91,92,93,94] and transition metal dichalcogenides (TMDs) [95].
MXenes are 2D materials from the transition metal carbide or nitride family with the general chemical formula Mn+1XnTx (M: early transition metal, X: carbon or nitrogen, Tx: surface functional groups, n = 1–4) [96]. MXenes simultaneously possess metal-level high electrical conductivity, hydrophilic surface characteristics, and excellent mechanical properties, and therefore have great potential as hybrid partners for Si anodes [96]. MXene sheets, which are either wrapped around Si nanoparticles or are inserted between these particles, play a role in mitigating the volume expansion and providing rapid electron transfer pathways [97]. Additionally, the abundant functional groups on the surfaces of MXenes can contribute to enhancing the interfacial stability by inducing strong interactions with both silicon and the electrolyte [98]. For example, hybrid structures that self-assembled as a result of the electrostatic attraction between Ti3C2Tx MXene sheets and Si nanoparticles demonstrated high capacity and excellent cycle stability [98]. Figure 7 details this electrostatic self-assembly approach. The schematic in Figure 7a illustrates how positively charged Si nanoparticles (nmSi-NH2) and negatively charged MXene sheets are mixed to form a self-assembled composite (BA electrode), which is then annealed. The performance graph in Figure 7b compares this BA electrode to a blank sample and an annealed-only (AA) electrode. The BA electrode exhibits significantly higher capacity retention over 100 cycles, confirming that the strong electrostatic interaction creates a uniform and stable coating of MXene nanosheets around the silicon, which effectively buffers volume expansion and provides a stable conductive network.
Hexagonal boron nitride (h-BN) has a 2D structure similar to that of graphene but is an electrical insulator with a wide bandgap [99]. BN is characterized by high thermal and chemical stability and excellent mechanical strength. Research in which BN was applied to silicon anodes primarily focused on exploiting the insulating properties and stability of BN to protect the silicon surface and control SEI formation [100]. Coating silicon nanoparticles with thin BN layers can help inhibit side reactions and form stable interfaces by preventing direct contact with the electrolyte. However, BN may, because of its inherently insulating nature, hinder electron transfer, thus additional compositing with other conductive materials such as carbon coatings is often necessary to ensure electrical conductivity [101].
TMDs are layered 2D materials with the chemical formula MX2 (M: transition metal, X: chalcogen element), with MoS2, WS2, and MoSe2 being representative examples [102]. TMDs exhibit semiconducting or metallic characteristics depending on their type and also possess lithium ion storage capabilities. Upon hybridization with Si, the TMD layers can act as buffer layers to mitigate the volume expansion experienced by Si, provide additional lithium storage sites, and improve the charge transfer characteristics at interfaces [103]. For example, the anode material consisting of silicon nanoparticles coated with MoS2 nanosheets improved the cycle stability and rate characteristics [101]. However, the electrical conductivity of TMDs themselves is lower than that of graphene or MXenes, and some TMDs may exhibit structural instability during reactions with lithium, which would necessitate further research. Table 3 provides performance data for hybrid systems combining silicon with 2D nanomaterials such as MXenes and MoS2. The data illustrates how the excellent mechanical flexibility and electrical conductivity of 2D materials contribute to improved cycling stability and rate performance of silicon anodes.
The use of 2D materials other than graphene also comes with its own set of challenges. Many transition metal dichalcogenides (TMDs) like MoS2 and WS2 are semiconductors with relatively low electrical conductivity, which can negatively impact the rate performance of the electrode. The synthesis of high-quality, large-area 2D materials can be complex and costly [104].
Furthermore, the interaction between the 2D material and the silicon surface needs to be carefully engineered to ensure good adhesion and efficient charge transfer. The long-term stability of these 2D materials in the electrochemical environment of a lithium-ion battery is also a concern, as they can undergo side reactions with the electrolyte [105]. The high cost and scalability issues associated with the production of many of these novel 2D materials are also significant barriers to their practical application.

3.1.4. Silicon–Conductive Polymer Hybrid Systems

Conductive polymers (CPs) are emerging as attractive hybrid partners for improving the performance of silicon anodes due to their unique electrical conductivity, flexibility, processing ease, and lightweight characteristics [106]. CPs can effectively encapsulate silicon particles or form three-dimensional network structures, which serve as flexible matrices that mitigate the volume expansion of Si [107]. Additionally, they provide electron and ion transfer pathways via their polymer chains, inhibit side reactions between silicon surfaces and electrolytes, and contribute to enhancing the mechanical stability of electrodes [108]. Representative conductive polymers that have been widely applied to hybrid systems with Si include polyaniline (PANI) [109], polypyrrole (PPy) [110], and poly (3,4-ethylenedioxythiophene):polystyrene sulfonate (PEDOT:PSS) [111].
Polyaniline (PANI) is a representative conductive polymer that is easy to synthesize, environmentally stable, and its electrical conductivity can be adjusted by modifying its doping state [112]. The following hybrid structures thereof have been researched: directly polymerizing and coating PANI on the surfaces of Si nanoparticles or dispersing Si particles within PANI matrices [113]. PANI coating layers, with their flexible characteristics, effectively accommodate the volume changes in Si, inhibit crack formation, and help maintain the structural integrity of electrodes [114]. Furthermore, PANI can control the reactions with electrolytes to induce stable SEI formation [115]. For example, hybrid electrodes comprising uniformly thin PANI layers coated onto silicon nanoparticles demonstrated excellent cycle stability and improved rate characteristics [116].
Polypyrrole (PPy) is also a conductive polymer that is relatively simple to synthesize, has excellent biocompatibility, and possesses good electrical conductivity and environmental stability [117]. PPy can be easily coated onto Si surfaces through chemical or electrochemical polymerization and serves as a protective layer that encapsulates the Si particles [118]. PPy coatings demonstrate effects that mitigate the volume expansion of Si, improve the electrical connectivity between particles, and inhibit electrolyte decomposition, thereby enhancing the lifespan of the electrode [118]. In particular, the application of a PPy coating to Si electrodes with three-dimensional porous structures reportedly simultaneously improves the structural stability and electrochemical performance [119]. Figure 8 showcases a specific example of this strategy. The schematic in Figure 8a illustrates the preparation of porous hollow Si (PHSi) spheres, which are subsequently coated with a polypyrrole (PPy) layer via in situ polymerization. The cycling data in Figure 8b highlights the dramatic effect of the PPy coating. The coated PPy@PHSi nanocomposite maintains excellent capacity and stability for over 250 cycles, whereas the uncoated version would typically degrade much faster. This result confirms that the flexible and conductive PPy layer successfully accommodates the volume changes of silicon and prevents the pulverization of the porous structure.
PEDOT:PSS, a material known to exhibit the highest electrical conductivity among conductive polymers developed to date, has the advantage of excellent processability as it is provided in the form of an aqueous dispersion [120]. PEDOT:PSS can be used to form thin and uniform coating layers on Si nanoparticle surfaces or to build conductive networks to connect Si particles to each other. The outstanding electrical conductivity of PEDOT:PSS is effective for improving the high-rate charge–discharge characteristics by greatly enhancing the electron transfer within electrodes. Additionally, the flexible structure of PEDOT contributes to buffering the volume changes in Si and enhancing the mechanical stability of electrodes [121]. Numerous studies have reported an improvement in the performance of the Si electrode by utilizing PEDOT:PSS as both a binder and conductive additive [122].
Silicon–conductive polymer hybrid systems are able to improve the performance in various ways depending on the type and structure of the conductive polymer and the method that was used to synthesize the polymer. PANI and PPy are relatively easy to synthesize and have excellent stability, while PEDOT:PSS provides outstanding electrical conductivity and processability. These hybrid systems generally exhibit improved cycle life and rate characteristics, with the flexibility of conductive polymers playing a particularly important role in mitigating the volume expansion problems associated with Si [123]. More recently, the focus has been on developing smart responsive polymer systems that actively control the volume expansion by changing their structure or properties in response to external stimuli (pH, temperature, voltage, etc.) beyond simple coating structures or on introducing polymer composites with self-healing functions that repair themselves upon being mechanically damaged to maximize the electrode durability [124]. Additionally, attempts to simultaneously optimize the electrical conductivity and mechanical flexibility by forming ternary hybrid systems by combining CPs with carbon materials are also being actively pursued.
Conductive polymers, despite their advantages, also have limitations. Their electrical conductivity is generally lower than that of carbon-based materials, which can limit the high-rate performance of the silicon anode. The long-term stability of conductive polymers during repeated charge–discharge cycles can be an issue, as they can degrade or lose their conductivity over time [125].
The synthesis and processing of conductive polymers can also be challenging. Achieving a uniform and conformal coating on silicon nanoparticles can be difficult, and the adhesion between the polymer and the silicon surface can be weak. Furthermore, the swelling of the polymer in the electrolyte can lead to mechanical instability and dimensional changes in the electrode [125]. The cost and scalability of the polymerization process are also important considerations for commercial applications. Table 4 summarizes the electrochemical performance of hybrid systems combining silicon with conductive polymers including PANI, PPy, and PEDOT:PSS. The quantitative data confirms that the flexible nature of conductive polymers effectively buffers silicon volume changes, thereby enhancing the mechanical stability of electrodes.

3.2. Classification According to Structural Characteristics

Structural design critically determines how effectively hybrid systems address silicon’s volume expansion challenges. Core–shell structures provide direct volume containment but face scalability challenges due to complex synthesis requirements. Sandwich structures [126] offer simpler fabrication but may compromise uniform stress distribution. Three-dimensional networks [127] excel in mechanical stability and conductivity but require precise control of porosity and interconnectivity for optimal performance. This section examines the design principles, advantages and disadvantages, performance enhancement mechanisms, and latest research trends of hybrid systems according to each of these three structural types.

3.2.1. Core–Shell Hybrid Structures

The core–shell structure—a structural form in which one material (the core) is completely encapsulated by another material (the shell)—is one of the most widely researched structures for solving the volume expansion problem of silicon anodes [34]. Generally, silicon is used as the core material, and is encapsulated in a shell consisting of a material that can buffer the volume expansion, impart electrical conductivity, or enhance the interfacial stability (carbon, MOx, CPs, etc.) to protect the silicon and improve the performance [128]. Core–shell structures can be either single- [129] or multi-shell structures [130] depending on the number of shells.
The single-shell core–shell structure is the most basic form in which the silicon core is encapsulated within a single layer of material. The shell material disperses the mechanical stress that occurs during the volume expansion of Si and prevents direct contact between the silicon surface and the electrolyte to inhibit the excessive formation of SEI layers [131]. For example, Si@C structures formed by encapsulating silicon nanoparticle cores within carbon shells exhibit improved cycle stability and rate characteristics owing to the excellent electrical conductivity and mechanical buffering effects of the carbon shells [129]. However, single-shell structures have limitations in that the shell may rupture or the shell may separate from the core during repeated extreme volume expansion.
To compensate for these disadvantages, multi-shell core–shell structures have been proposed. Multi-shell structures are formed by sequentially forming shell layers composed of two or more different materials around a silicon core [132]. Each shell layer can be designed to perform specific functions. For example, a representative structure (Si@void@Shell) contains an inner shell to create void space directly around the silicon core, and an outer shell that is mechanically robust or has excellent electrical conductivity on the outside [132]. The internal void space provides sufficient room for the volume expansion of Si and minimizes the stress impaired to the outer shell, while the outer shell maintains the stability of the entire structure and provides electrical conduction pathways [45]. These multi-shell structures have superior volume expansion mitigation effects compared to single-shell structures and are advantageous for securing long-term cycle stability. For example, various multi-shell structures such as Si@void@C [133], Si@SiOx@C [134], and Si@TiO2@C [135] have been reported and have demonstrated excellent electrochemical performance. Figure 9 illustrates an advanced core–shell design, the Si@void@C structure, and its superior performance. The process is shown in Figure 9a, where a dissolvable resin layer is used to create a precise void space between the Si core and the outer carbon shell. The TEM images in Figure 9b visually confirm the successful formation of this yolk–shell architecture. The cycling performance in Figure 9c is particularly impressive, showing that the Si@void@C anodes maintain stable cycling for over 1000 cycles at a very high rate of 2.0 A/g, far outperforming bare Si. This demonstrates conclusively that the internal void is the key to accommodating volume expansion without stressing and fracturing the protective outer shell.
The use of a core–shell structure is a highly effective strategy for improving the electrochemical performance by effectively controlling the volume expansion of Si and enhancing the interfacial stability. The performance can be optimized by precisely controlling the type, thickness, and number of shell materials, as well as the interfacial characteristics between cores and shells. Recent research has focused on developing functional multi-shell structures that impart specific functions to each shell layer (e.g., selective lithium ion permeation, self-healing functions) beyond simple multi-shells [136], or designing hierarchical core–shell systems where nanosized core–shell units are interconnected to form macroscopic hierarchical structures [137]. These advanced core–shell structural designs are expected to make important contributions to overcoming the performance limitations of silicon anodes to increase their commercialization potential. For instance, a multi-level confinement strategy, where porous silicon is wrapped in graphene oxide and carbon nanotubes and then coated with carbon (pSi@GO-CNTs@C), has been shown to effectively buffer volume expansion and significantly enhance long-term cycling stability [138].

3.2.2. Sandwich Hybrid Structures

Sandwich structures refer to forms in which different material layers are alternately stacked such as in a sandwich, and are attracting attention as effective structures particularly when combining 2D nanomaterials with silicon [139]. In these structures, the silicon layer primarily provides high capacity, while the other material layers (e.g., graphene, MXenes, or other 2D materials or CPs) serve as buffering layers to accommodate the volume expansion of Si, electron conduction pathways, or mechanical support structures [140]. Sandwich structures can be classified into single- [141] and multilayer sandwich structures [50] according to the number of layers.
Single-layer sandwich structures generally take the form of inserting a silicon layer between two layers consisting of another material (e.g., graphene/silicon/graphene). The outer layers constrain the volume expansion experienced by Si from both sides, maintain the structural stability of the entire electrode, and provide excellent electrical conduction pathways [141]. Particularly when using a 2D material such as graphene for the outer layers, their large specific surface area and flexibility effectively accommodate the volume changes in Si while preventing particle agglomeration and enabling rapid electron transfer. For example, structures comprising a layer of silicon nanoparticles inserted between two graphene sheets have been reported to exhibit high reversible capacity and excellent cycle stability [142]. However, an excessively thick silicon layer may impede the diffusion of lithium ions to the interior of the silicon layer, or the structure may be destroyed as the stress applied to the outer layers increases during volume expansion, which is a disadvantage.
Attempts to solve these problems have inspired research on multilayer sandwich structures. These structures consist of layers of silicon and other materials (buffering/conducting layers) alternately stacked in multiple layers. By controlling the thickness of each layer at the nanometer level, the migration distance of lithium ions and electrons can be shortened, and the stress applied to the entire structure can be minimized by dispersing the volume expansion of Si across each layer [143]. For example, multilayer structures created by alternately depositing thin silicon films and graphene layers have shown very high rate capabilities and long-term cycle stability [144]. Additionally, attempts are being made to facilitate ion transfer and more effectively mitigate the volume expansion by introducing void spaces between layers or inserting functional materials (e.g., ion-conductive polymers) [145]. Figure 10 provides a representative example of the sandwich structure approach. The fabrication process is detailed in Figure 10a, where layers of Si nanoparticles and reduced graphene oxide (RGO) are alternately deposited on a Ni foam substrate. The schematic in Figure 10b depicts how the flexible RGO layers encapsulate the Si nanoparticles, constraining their volume expansion during lithiation. The electrochemical data in Figure 10c confirms the effectiveness of this design; the multilayered Si/RGO anode shows significantly higher capacity retention and stability at a 1C rate compared to a simple Si anode, proving the dual function of RGO as both a mechanical buffer and a conductive pathway.
Sandwich structures are promising strategies for enhancing the performance of silicon anodes, particularly by maximizing the advantages of 2D materials. Layered structures effectively constrain volume expansion, provide excellent electron and ion transfer pathways, and contribute to increasing the mechanical stability of electrodes. Customized performance implementation is possible by precisely designing the structure by specifying the type, thickness, number of layers, and interfacial characteristics of the materials between layers. Recent research has focused on introducing functional interlayer materials that impart specific functions (e.g., self-healing, selective ion permeation) between layers beyond simple alternating stacking structures, or implementing more complex and multifunctional sandwich structures by utilizing nanostructures of other dimensions such as nanowires and nanotubes alongside 2D materials [146]. These advanced sandwich structure designs are expected to play an important role in simultaneously enhancing the energy density and stability of silicon anodes.

3.2.3. Three-Dimensional Network Hybrid Structures

Three-dimensional network (3D network) structures refer to structures where silicon active materials and other functional materials (e.g., CNTs, graphene, metal nanowires, etc.) are interconnected to form macroscopic 3D frameworks [147]. Unlike individual particle-level structures such as core–shell or sandwich structures, these 3D network structures offer several advantages by providing interconnected porous structures across the entire electrode. Their large specific surface area broadens the contact interface with the electrolyte, enhances their lithium ion accessibility, and the interconnected networks provide efficient electron transfer pathways throughout the electrode [148]. Additionally, existing pores within the network provide space for the volume expansion of Si and contribute to increasing the structural stability of electrodes by effectively mitigating the mechanical stress that accompanies charge–discharge cycling [148]. These 3D network structures can be classified into porous 3D networks [149] and hierarchical 3D networks [150] according to the pore size and distribution, as well as the hierarchy of the network framework. Figure 11 demonstrates an example of a hierarchical 3D network structure. The schematic in Figure 11a shows the self-assembly of Si nanoparticles and CNTs into a 3D hydrogel, which is then freeze-dried and annealed to create a CNTs-crosslinked hierarchical porous Si (CHSP) structure. The rate performance graph in Figure 11b compares this CHSP anode with a standard nano-Si anode. The CHSP structure exhibits vastly superior capacity at all current densities from 0.1 to 5.0 A/g, highlighting that the interconnected porous 3D network provides excellent pathways for both ion and electron transport while also accommodating volume expansion, which is critical for high-rate performance.
Simple porous 3D network structures consist of silicon active materials dispersed within a framework in which pores with a relatively uniform size are three-dimensionally connected. These structures can be fabricated by various methods such as electrospinning [151], freeze-drying [152], and template removal methods [153]. Examples are methods that involve impregnating a 3D non-woven matrix (formed by electrospinning CNFs) with silicon nanoparticles [154], or depositing Si onto porous carbon structures such as graphene aerogels [152]. Porous networks provide expanded surface areas and interconnected conductive frameworks, which enable high capacity and excellent rate characteristics [155]. The pore structure effectively accommodates the volume expansion of Si to enhance the cycle stability. However, the disadvantages of simple porous structures are that stress may concentrate in specific areas or ion/electron transfer pathways may be inefficient if the pore size or distribution is non-uniform [156].
Efforts to overcome these limitations have led to the proposal of hierarchical 3D network structures. Hierarchical structures refer to complex network structures, in which pores of various sizes ranging from micrometer-sized macropores to nanometer-sized micropores/mesopores, are hierarchically connected [157]. Macropores facilitate electrolyte penetration and provide pathways for rapid ion movement, while micropores maximize the contact area with active materials and effectively buffer the volume expansion [158]. Furthermore, hierarchical network frameworks can further enhance the mechanical stability. These hierarchical structures can be fabricated by mimicking biological structures (e.g., trees, sponges) or by forming precisely controlled structures using self-assembly phenomena. For example, hybrid electrodes with silicon nanoparticles uniformly dispersed within hierarchical porous carbon structures have been reported to simultaneously achieve outstanding rate characteristics and long-term cycle stability [63].
Hybrid structures consisting of a 3D network offer an effective approach for solving the volume expansion problems associated with Si at the electrode level and maximizing the electrochemical performance. Interconnected porous structures facilitate smooth ion and electron transfer, provide excellent buffering capacity for volume changes, and enhance the mechanical strength of electrodes [159]. In particular, hierarchical 3D network structures present the possibility of further optimizing the performance by including pores of various sizes and precisely controlled network frameworks. Recent research has focused on the design of biomimetic 3D structures that mimic efficient structures in nature [160], or on the development of self-assembly 3D network technologies that spontaneously form the desired structures via molecular-level interactions [148]. Additionally, research using 3D printing technology to directly fabricate customized 3D electrode structures is also actively underway [161]. These innovative approaches to design 3D network structures are expected to establish themselves as key strategies for developing high-performance, highly stable silicon anodes.

4. Synthesis Methods for Hybrid Material Systems

The performance of hybrid material systems for silicon anodes depends greatly not only on the types of constituent materials and structures but also on the synthesis methods used to process them. The synthesis method is a key element that determines the uniformity of hybrid structures, interfacial characteristics between components, and the physicochemical properties of the final material. Therefore, selecting appropriate synthesis methods and optimizing the process parameters are crucial for obtaining hybrid systems with the desired structures and performance. The synthesis of hybrid material systems can be broadly divided into solution-based synthesis methods [30], vapor deposition methods [30], mechanical synthesis methods [162], and hybrid processes combining these methods [163].

4.1. Solution-Based Synthesis Methods

Solution-based synthesis methods involve the dissolution or dispersion of the precursor materials in solvents, followed by the synthesis of the desired hybrid materials through chemical reactions (hydrolysis, condensation, precipitation, etc.) [160]. These methods offer advantages such as processing at relatively low temperatures, suitability for mass production, and the ability to produce materials in various forms (powders, thin films, fibers, etc.). Additionally, the composition, size, shape, and crystallinity of the final products can be easily controlled by adjusting the chemical reaction conditions of the solution [164]. Solution-based synthesis methods that are primarily used for fabricating hybrid systems for silicon anodes include the sol–gel method [165], hydrothermal synthesis [166], and the microemulsion method [167].
The sol–gel method involves hydrolyzing and condensing precursors such as metal alkoxides or salts in solvents (mainly water or alcohol) to form a sol state, which is subsequently converted to a gel state through additional reactions or drying processes, and then heated to obtain the final oxides or composites [168]. This method offers advantages such as that it enables uniform mixing at the molecular level for easy composition control, is favorable for forming porous structures, and enables processing at relatively low temperatures [169]. In the field of silicon anodes, it is utilized to manufacture silicon–oxide or silicon–carbon hybrid structures via sol–gel processes with silicon precursors and other metal oxides (e.g., TiO2, SiO2) or carbon precursors [170]; for example, the dispersion of silicon nanoparticles within porous TiO2 matrices manufactured by the sol–gel method, or the formation of uniform SiO2 or carbon coating layers on silicon particle surfaces [170]. Disadvantages of the sol–gel method include long processing times, the possibility of crack formation due to shrinkage during drying, and cost and environmental issues associated with organic solvents or precursor materials. Figure 12 details the synthesis of a Si@HC (Hard Carbon) yolk–shell structure using a sol–gel based method. As shown in Figure 12a, the process begins with coating Si nanoparticles with a SiO2 and resin shell, followed by carbonization and HF etching to create the final void-containing structure. The TEM images in Figure 12b clearly visualize the formation of the void space inside the carbon shell. The electrochemical data in Figure 12c shows that the resulting anodes exhibit stable cycling and good rate capability, demonstrating that the sol–gel method is a versatile tool for fabricating complex, well-defined nanostructures for battery applications.
Hydrothermal synthesis entails the use of water or other solvents as reaction media under high-temperature and high-pressure conditions to synthesize materials [171]. Because these reaction conditions increase the solubility of reactants and the reaction rates, they are advantageous for synthesizing highly crystalline materials or materials with unique microstructures that are difficult to obtain through conventional solution reactions. Hydrothermal synthesis can be accomplished with relatively simple equipment, can produce nanoparticles of uniform size and shape, and is widely used for synthesizing various types of oxides, sulfides, metals, and other materials [171]. In the field of silicon anodes, it is applied to directly grow metal oxide (e.g., SnO2, Fe3O4) nanostructures on the surfaces of silicon nanoparticles [172] or to form hybrid composites by hydrothermally treating silicon with other materials (e.g., graphene oxide) [173]. For example, graphene-wrapped silicon (Si@G) structures that were synthesized by hydrothermally reacting silicon nanoparticles with graphene oxide demonstrated enhanced electrical conductivity together with the ability to mitigate volume expansion effects [174]. Disadvantages of hydrothermal synthesis are the requirement for high-pressure reactors, which requires safety precautions to be observed, and the difficulty to precisely control reaction conditions.
The microemulsion method involves forming thermodynamically stable micro- or nanometer-sized droplets by adding surfactants to two immiscible liquid phases such as water and oil, and using these droplets as nanoreactors to synthesize materials [175]. The droplet size in microemulsions can be controlled by adjusting the type and ratio of surfactants and solvents, with the advantage of enabling the size and shape of the resulting particles to be precisely controlled. This method is effective for synthesizing nanoparticles of uniform size or fabricating complex nanostructures such as core–shell structures [175]. In silicon anode research, it is utilized to reduce the silicon precursors within microemulsion droplets to produce silicon nanoparticles of uniform size [176], or to create core–shell structures by using silicon nanoparticles as cores and forming shells of other materials (e.g., carbon, metal oxides) on their surfaces using microemulsions [177]. To its disadvantage, the microemulsion method requires relatively complex system configurations, may incur cost and environmental issues due to the use of large amounts of surfactants and organic solvents, and may be difficult to scale for mass production.
In solution-based synthesis methods, various synthesis conditions such as the type and concentration of precursors, type of solvent, reaction temperature and time, pH, and additives greatly influence the structure and physical properties of the final hybrid material. Therefore, systematically optimizing these process variables is essential to achieve the desired performance. Recent research has focused on eliminating the disadvantages of existing solution-based synthesis methods; for example, the development of synthesis methods that use environmentally friendly solvents such as water or bio-based solvents to minimize damage to the environment [178], the development of low-temperature process technologies to lower the energy consumption [179], and the utilization of microwaves or ultrasound as auxiliary means to shorten reaction times and increase the uniformity [180]. Additionally, attempts to synthesize more complex and sophisticated hybrid structures by combining multiple solution-based processes or integrating with other synthesis methods are ongoing [181].

4.2. Vapor Deposition Methods

Vapor deposition provides a solvent-free and vacuum-based way to form high-purity, conformal, and precisely thickness-controlled coatings on complex silicon architectures, which helps buffer volume change and stabilize the SEI, and the following sections will detail how conductive carbon shells grown by CVD and angstrom-level inorganic layers deposited by ALD translate these advantages into durable, high-rate silicon anodes.
CVD entails the deposition of the desired materials in thin film form through chemical reactions on a substrate surface or in the gas phase by injecting gaseous precursors into a reaction chamber [182,183]. The composition, structure, and thickness of the thin film can be controlled by adjusting the reaction temperature, pressure, precursor flow rate, and ratio. CVD offers advantages such as uniform thin film deposition over relatively large areas, applicability to substrates with complex shapes, and the ability to deposit various types of materials (metals, semiconductors, insulators, etc.) [184]. In the field of silicon anodes, CVD is primarily utilized to deposit carbon coatings (e.g., a layer of pyrolytic carbon) on silicon nanostructures or to directly grow carbon-based 2D materials such as graphene [185]. Examples are the synthesis of silicon thin films or nanowires by thermally decomposing silane (SiH4) gas [186], or the deposition of carbon layers on silicon surfaces by decomposing hydrocarbon gases such as methane (CH4) or acetylene (C2H2) at high temperatures [187]. CVD processes generally require high temperatures, and the precursor gases used are often toxic or explosive. This necessitates the use of safety equipment, which is a disadvantage.
Atomic layer deposition (ALD) is a type of CVD for the deposition of thin films consisting of atomic layers on a substrate surface via self-limiting chemical reactions by sequentially injecting the precursors and reactants in pulse form [188]. Because further reactions cannot occur once the surface reactions are saturated at each pulse step, the film thickness can be very precisely controlled at the atomic layer level by adjusting the number of pulses. ALD has the unique advantage of being able to deposit extremely thin, uniform, and pinhole-free high-quality thin films with high conformality even on complex 3D nanostructure surfaces [189]. In silicon anode research, ALD is actively applied to form very thin protective layers (e.g., Al2O3, TiO2, or ZnO) on the surfaces of silicon nanoparticles to enhance the SEI stability [189], or to deposit metal coatings or conductive oxide thin films for improved conductivity [190]. For example, the uniform deposition of Al2O3 layers with a thickness of a few nanometers on the surfaces of silicon nanoparticles using ALD can effectively inhibit the unwanted volume expansion and greatly improve the cycle life [191]. The disadvantages of ALD are its very slow deposition rates, requirement for expensive equipment, and potentially limited selection of suitable precursors and reactants.
Sputtering is a representative PVD method that generates the ionized plasma of an inert gas, such as argon (Ar), in a vacuum chamber and accelerates these ions to collide with the surface of a target material at high energy. This causes the target atoms to be ejected and deposited onto the substrate surface [192]. Sputtering can be used to deposit almost all types of metals, alloys, and ceramic materials in the form of a thin film, can be used at relatively low temperatures, and produces thin films with excellent adhesion [192]. In the field of silicon anodes, sputtering is used to directly deposit silicon thin films or to deposit metal coatings (e.g., Cu, Ni) or alloy thin films that serve as conductivity enhancement or protective layers on silicon nanostructures. For example, electrodes have been fabricated by depositing silicon thin films on porous current collectors using sputtering [193], or the electrical conductivity and structural stability have been improved by coating thin copper (Cu) layers on the surfaces of silicon particles [194]. Among the disadvantages of sputtering processes are their relatively low deposition rates, lower conformality for substrates with complex shapes compared to CVD or ALD, and the possibility of damaging the substrate with the plasma. Figure 13 illustrates the use of sputtering to improve Si anode performance. The schematic in Figure 13a shows the deposition of a Cu layer onto Si particles via magnetron sputtering. The electrochemical performance is shown in Figure 13b,c. The results clearly indicate that both the cycling stability and rate capability of the Si anode are significantly improved after being coated with Cu (Si@Cu) and further with carbon (Si@Cu@C). This confirms that a sputtered metallic layer can effectively enhance the electrical conductivity and provide some mechanical support, thereby mitigating some of the key issues of Si anodes.

4.3. Mechanical Synthesis Methods

Mechanical synthesis methods involve synthesizing the desired hybrid materials by applying mechanical energy to raw materials in the solid state by way of grinding or mixing to induce alloying or chemical reactions [162]. Because these methods are dry processes that do not use solvents, they offer advantages such as the possibility of mass production with relatively simple equipment and applicability to combinations of various types of raw materials [162]. Particularly, they can be usefully employed to process materials that cannot tolerate high temperatures, as they can be processed at room temperature. Ball milling, also known as high-energy ball milling (HEBM), is a representative mechanical synthesis method.
In the field of silicon anodes, ball milling is widely used to manufacture hybrid composites of silicon with carbon-based materials (graphite, CNTs, graphene, etc.) or other metals/metal oxides. For example, when micrometer-sized silicon powder is ball milled with graphite powder, the silicon particles are finely ground while they are simultaneously uniformly dispersed within the graphite particles or attached to the surfaces of the graphite particles to yield silicon–graphite composites [195]. These composites can mitigate the low conductivity and volume expansion problems of Si by utilizing the excellent electrical conductivity and structural stability of graphite [195]. Research on manufacturing hybrid structures comprising silicon particles surrounded by CNT networks or graphene sheets by ball milling silicon with CNTs or graphene is also being actively conducted [196]. The mechanical forces generated during the ball milling process can induce strong bonding between silicon and the carbon materials, which contributes to lowering the interfacial resistance and enhancing the structural stability [197]. Figure 14 demonstrates the effectiveness of the ball-milling process for creating Si/CNT composites. The cycling and rate performance graphs in Figure 14a,b show that the Si30h/CNT composite, prepared by ball-milling, has vastly superior stability and rate capability compared to the Si30h material alone. The Nyquist plot in Figure 14c reveals a much lower charge-transfer resistance for the composite, which is direct evidence of the improved electrical network formed by the CNTs. The SEM/TEM images in Figure 14d visually confirm that the CNTs are well-distributed and entangled with the Si particles, providing the structural basis for the performance enhancement.
Ball milling is a relatively simple and cost-effective process suitable for mass production, but it also has several disadvantages. First, the high energy generated during the ball milling process may degrade or amorphize the crystallinity of the raw materials, which can affect the electrochemical performance [198]. Second, impurities originating from milling containers or balls may be incorporated into the final product, which can be problematic especially in battery materials that require high purity [199]. The selection of the container and ball materials and process management are therefore important. Third, particles manufactured by ball milling may have non-uniform sizes or shapes, and agglomeration can easily occur, and may potentially require additional dispersion processes [200].
Research being conducted to overcome these disadvantages to improve the efficiency and controllability of ball milling processes includes, for example, methods to prevent excessive cold welding or agglomeration of particles and increase the grinding efficiency by adding small amounts of liquid (the process control agent, PCA) during milling [201]. Additionally, attempts to induce specific chemical reactions or control the crystallinity by regulating the milling temperature or atmosphere have been reported [202]. Other researchers have combined ball milling with other processes (e.g., heat treatment, chemical etching, etc.) to produce more complex and functionalized hybrid structures; for example, porous silicon structures were produced by removing the metal components through selective etching after manufacturing silicon–metal composites by ball milling [203], or by restoring the crystallinity or forming specific phases through heat treatment after ball milling [24]. Mechanical synthesis methods, especially ball milling, are important synthesis technologies with the potential to mass-produce silicon-based hybrid anodes, and their range of applications is expected to expand further through process optimization and integration with other technologies.

4.4. Hybrid Process Methods

Hybrid process methods involve combining two or more of the previously described solution-based synthesis methods, vapor deposition methods, and mechanical synthesis methods [204]. These methods were designed to manufacture hybrid material systems with complex structures or characteristics that are difficult to implement with a single process alone. By taking the advantages of each process and compensating for their disadvantages, hybrid materials for silicon anodes with more precisely controlled structures and excellent performance can be developed. Hybrid processes can be designed in various ways depending on the types and sequences of the processes that are combined, which significantly influence the structure and characteristics of the final material.
One of the most common hybrid processes combines solution-based synthesis methods with vapor deposition methods. For example, support structures with porous structures (e.g., carbon aerogels, metal oxide nanostructures) can first be fabricated using solution processes such as sol–gel or hydrothermal synthesis, after which silicon thin films or nanostructures can be deposited on these support structures using vapor deposition methods such as CVD or sputtering [205]. This approach can simultaneously exploit the advantages of 3D porous structures obtained through solution processes (large surface area, volume expansion mitigation space) and the advantages of high-quality silicon thin films deposited using vapor deposition methods (uniformity, high initial capacity) [205]. Conversely, hybrid processes to first synthesize silicon nanostructures (e.g., nanowires, nanotubes) using vapor deposition methods, followed by the formation of protective layers or conductivity enhancement layers on their surfaces using ALD or solution-based coating methods (e.g., sol–gel coating, polymer coating) are also widely used [191]. For example, coating extremely thin and uniform Al2O3 protective layers on the surfaces of silicon nanowires grown by CVD using ALD can greatly enhance the electrochemical performance and stability [206].
Hybrid approaches in which mechanical synthesis methods are combined with other processes are also effective. For example, silicon can be mechanically mixed with carbon materials using ball milling to produce composite powders, which are then calcinated to form structures in which the silicon particles are dispersed within a carbon matrix or to enhance the crystallinity [163]. In this case, ball milling is responsible for uniform mixing and particle size reduction, whereas the heat treatment plays a role in controlling carbonization, crystallization, or phase formation. Additionally, spherical secondary particles or fiber-form electrode active materials were manufactured by combining silicon-based composite powders by ball milling with solution-based processes such as spray drying or electrospinning [207]. These hierarchical structures can enhance the ease of handling powders and improve the electrode manufacturing processability, while simultaneously enhancing the electrochemical performance by controlling the micro-/nanostructure [208].
Recently, more complex hybrid processes that combine three or more processes in sequence or in parallel have been developed. For example, highly precisely controlled multi-shell or multicore structures can be implemented through multistage hybrid processes such as creating porous carbon supports using solution processes, depositing silicon on them using vapor deposition methods, and then performing surface modification using solution coating or ALD [209]. These complex structures, which are designed to maximize the functions of each component and rely on synergistic effects, enable performance improvements at levels that were difficult to realize with existing single or dual structures.
Hybrid process methods provide powerful tools to overcome the performance limitations of silicon anodes by integrating the advantages of various synthesis technologies. However, disadvantages such as process variables becoming complicated when combining multiple processes, the importance of compatibility and optimization between each stage, and the potential increase in overall process costs must also be considered. Therefore, an important aspect of designing effective hybrid processes is to optimize the process combinations and sequences based on a deep understanding of each unit process, along with clear designs for the final target structures and performance. Future research is expected to proceed in directions that enhance precise control technologies for each process stage, maximize synergistic effects between processes, and increase the efficiency and economic feasibility of overall processes. Notably, attempts to efficiently explore vast process variable spaces and predict optimal hybrid process conditions using artificial intelligence (AI) or machine learning technologies are also attracting attention [210].

5. Advantages, Disadvantages, and Challenges of Hybrid Material Systems

The various hybrid material systems mentioned above have been developed to mitigate the chronic problem of volume expansion in silicon anodes to enhance their electrochemical performance. These hybrid approaches have greatly contributed to maintaining the advantageous high theoretical capacity of Si while complementing its disadvantages and imparting new functionalities through synergistic effects with heterogeneous materials such as carbon, metal oxides, 2D materials, and CPs. In this section, the main advantages, disadvantages, and limitations of hybrid material systems are examined, the challenges that need to be overcome for commercialization are discussed, and possible solutions are proposed.

5.1. Advantages of Hybrid Material Systems

The greatest advantage of hybrid material systems is that they can be designed to effectively control the volume expansion of Si. Carbon coating layers [211] or matrices [212], metal oxide shells [213,214], and graphene wrapping [215] provide physical buffer spaces around silicon particles or impart mechanical constraint forces to disperse the stress generated during charging and discharging to inhibit the destruction of the electrode structure. Figure 15 provides a clear and powerful demonstration of this principle. The graph compares the cycling performance of an uncoated Si anode with anodes coated with 5 nm and 10 nm of TiO2. While the uncoated anode’s capacity fades rapidly, especially at higher C-rates, both TiO2-coated anodes exhibit remarkable stability. This directly confirms that even an ultrathin metal oxide shell can act as an effective mechanical buffer, successfully mitigating volume expansion and drastically improving the cycle life of the electrode. The ability to control the volume expansion is of key importance for greatly improving the low cycle stability, the most serious problem presented by pure silicon anodes. The ability to control the volume expansion is of key importance for greatly improving the low cycle stability, the most serious problem presented by pure silicon anodes. Hybrid systems are also able to effectively compensate for the low electrical conductivity of Si. Compositing with conductive materials such as carbon or CPs secures electron transfer pathways to lower the internal resistance of electrodes, which leads to high rate capabilities and low voltage polarization [40]. Consequently, rapid charging and discharging become possible, which improves the energy efficiency. The ion transfer characteristics can also be enhanced through structural design. Various hybrid structures such as core–shell, sandwich, and 3D networks facilitate ion transfer by shortening the lithium ion diffusion distances or optimizing the contact areas between the active material and electrolyte [67]. This is particularly important in applications that require high power. Furthermore, certain hybrid partner materials provide additional functions. For example, some metal oxides or 2D materials can contribute to improving the battery life and safety by inducing stable SEI layer formation on silicon surfaces or by serving as protective layers that inhibit electrolyte decomposition [216,217]. In this way, hybrid material systems offer the potential to increase the electrochemical performance to vastly beyond that of pure silicon by complementing the disadvantages and maximizing the advantages of Si by combining various materials structural designs.

5.2. Disadvantages and Limitations of Hybrid Material Systems

Despite their many advantages, hybrid material systems have several inherent disadvantages and limitations. The most representative is the trade-off relationship between capacity and stability. Although inactive materials constituting hybrid systems such as carbon and metal oxides contribute to mitigating the volume expansion and increasing the conductivity of Si, they themselves have low or almost no lithium storage capacity [218]. Therefore, an increase in the amount of hybrid partner materials necessarily lowers the reversible capacity of the entire electrode [219]. This dilemma arises when the proportion of inactive materials is increased to enhance the electrode stability, but dilutes the advantage of the inherently high capacity of Si. A major disadvantage is that the synthesis of hybrid structures can be complex and costly. Particularly in cases requiring elaborate nanostructure control such as core–shell structures or multilayer structures, or hybrid process methods that require multiple sequential processes, low synthesis yields or high process costs can become obstacles to mass production and commercialization [220]. Even materials that deliver excellent performance at the laboratory scale may be difficult to apply in practice without economically viable mass production technologies. Interface issues within hybrid systems are also important limitations. Interfaces between silicon and hybrid partner materials play important roles in electron and ion transfer, stress dispersion, and other functions, but the interfacial resistance may be high or the bonding strength may be weak due to physical and chemical mismatches between the two materials. The occurrence of delamination or cracks at the interfaces during repeated charging and discharging may obstruct conduction pathways to cause the performance to rapidly deteriorate. The formation and maintenance of stable and efficient interfaces therefore remains a technical challenge. Lastly, the still low ICE is also a limitation [221]. Due to the large surface area of Si and the complexity of hybrid structures thereof, the amount of lithium consumed during SEI formation in the first charge is often high, with the result that the ICE is lower compared to that of graphite anodes [197]. Figure 16 provides a direct comparison that highlights this critical issue. The graph shows that while the Sn-Si hybrid composite anode delivers a higher specific capacity than the graphite anode, its first-cycle Coulombic efficiency is significantly lower (below 70%). This starkly contrasts with graphite’s efficiency of over 90%, visually confirming that the large irreversible capacity loss from SEI formation on high-surface-area hybrid materials is a major hurdle to overcome for practical applications. This, in turn, reduces the energy density of full cells and lowers the utilization of the cathode materials, a problem that still requires attention.

5.3. Challenges to Overcome and Possible Solutions

Continuous research and development are necessary to overcome the disadvantages and limitations mentioned above to ensure the successful commercialization of silicon anodes based on hybrid material systems. The main challenges and possible solutions are as follows. First, the materials and structural designs need to be optimally combined to simultaneously achieve high capacity and high stability. Rather than simply increasing the inactive material content, strategies are needed to develop new hybrid partner materials that provide effective buffering and conductivity enhancement effects even in small amounts, or to control the structure of Si (e.g., porous, nanowire) to minimize the volume expansion [222]. Currently, research is actively underway to simultaneously secure the structural stability and ion/electron transfer characteristics while maintaining high silicon content through 3D hierarchical structures or spherical secondary particle designs [64,223]. Second, low-cost, high-efficiency mass synthesis processes need to be developed urgently. Instead of complex laboratory-scale processes, the development of simple and scalable synthesis methods that would be industrially applicable is essential. For example, processes that include mechanical synthesis methods (ball milling) need to be optimized, solution-based processes should be simplified [224], the yield of vapor deposition methods needs to be enhanced, and the development of practical hybrid processes that combine the advantages of each process is also important [225,226]. Taking a long-term perspective, inexpensive raw materials need to be identified and recycling technologies should be developed. Third, interfacial engineering research to lower the interfacial resistance and strengthen the bonding is important. Attempts are being made to induce chemical bonding between silicon and hybrid partner materials or to improve the interfacial stability and ion/electron transfer characteristics by introducing ultrathin artificial interfacial layers using ALD and other methods [227,228]. The utilization of advanced analytical techniques to elucidate detailed reaction mechanisms at interfaces should also be conducted in parallel. Fourth, strategies to improve the ICE are needed. Technologies such as pre-passivation treatment of silicon surfaces, artificial SEI layer formation, or pre-lithiation techniques that add limited amounts of lithium in advance are being researched [229]. These technologies can contribute to enhancing the ICE by reducing irreversible lithium consumption in the first cycle and improving the full-cell performance. Figure 17 illustrates a material-level approach to this problem. By comparing the initial charge–discharge profiles of a Si/C anode and a Si/Graphite/C (Si/G/C) anode, the graph shows that the Si/G/C composite has a smaller irreversible capacity (the difference between the first charge and discharge capacities). This suggests that incorporating stable, low-surface-area graphite into the composite can mitigate excessive side reactions, thus serving as a practical strategy to improve the ICE.
Fifth, securing long-term reliability in actual operating environments and improving the electrode processability are necessary. The performance of the electrode in full cells should be optimized based on the half-cell evaluation results and comprehensive verification including high/low temperature characteristics, lifetime prediction, and safety assessment should be conducted [230]. Additionally, the development of technologies to control the powder properties (density, particle size, flowability, etc.) such that the hybrid materials that are developed are well compatible with existing electrode manufacturing processes (slurry preparation, coating, calendering, etc.) is also important [231]. Figure 18 showcases an advanced material design that considers these practical aspects. The schematic details a method to create a HPC/Si@C composite by embedding Si within a hierarchical porous carbon (HPC) host. This approach not only tackles the volume expansion issue at the particle level but also results in larger, more uniform secondary particles, which can improve powder flowability and slurry characteristics, making it more compatible with existing electrode manufacturing lines. Solving these challenges is expected to require the active utilization of computational science and data-based research methodologies, along with convergence research across various fields such as materials science, electrochemistry, and process engineering.
Figure 19 illustrates a particularly important strategy for commercialization: developing low-cost, scalable manufacturing processes. The schematic shows the fabrication of porous Si (P-Si) by simply etching aluminum from an inexpensive, metallurgical-grade Si-Al alloy, followed by a carbon coating. This method bypasses the need for costly Si nanoparticles as a starting material. By demonstrating a pathway to produce high-performance porous silicon anodes from abundant, low-cost raw materials, this approach directly addresses the economic barriers that often hinder the mass production of advanced battery materials.

5.4. Recent Industrial Progress and Commercial Cases

Despite the challenges, significant progress has been made in the commercialization of silicon anodes, with several companies leading the charge. This section delves into the specific strategies and achievements of key players in the industry.
(i) Tesla: As a trailblazer in the EV market, Tesla has been at the forefront of battery innovation. While initial reports suggested that their 4680 cells would feature a high-silicon anode, recent teardowns have revealed that the current generation of 4680 cells still relies on a graphite-based anode. This highlights the immense difficulty in achieving stable, long-lasting performance with high silicon content in a mass-produced cell. However, Tesla continues to invest heavily in R&D to overcome these hurdles. Their patents and research papers indicate a focus on developing novel electrode structures, such as silicon nanowires or porous silicon, combined with advanced binders and electrolyte additives to mitigate the volume expansion issue and improve cycle life. The company’s long-term goal is to gradually introduce more silicon into their anodes to further increase the energy density and driving range of their vehicles.
(ii) Sila Nanotechnologies: Sila is a prominent US-based company that has developed a silicon-dominant anode material called “Titan Silicon”. This material is a nano-composite silicon (NCS) powder that is designed to replace graphite in existing battery manufacturing processes. Sila’s approach involves creating a porous, engineered structure at the nanoscale that can accommodate the volume expansion of silicon without fracturing. The company claims that its material can increase the energy density of lithium-ion batteries by 20–40% compared to graphite-based anodes. Sila has already achieved commercial success in the consumer electronics market, with its material being used in the WHOOP 4.0 fitness tracker. The company is now focused on scaling up its production for the automotive market and has established a large-scale manufacturing plant in Moses Lake, Washington, which is expected to start production in 2025. Sila has also signed a supply agreement with Panasonic to provide its Titan Silicon material for EV batteries.
(iii) Panasonic: As a major battery manufacturer and a key partner of Tesla, Panasonic is actively pursuing the integration of silicon anodes into its products. The company has a two-pronged strategy. First, it is working on gradually increasing the silicon content in its existing graphite-based anodes. This incremental approach allows them to improve energy density while maintaining the reliability and safety of their batteries. Second, Panasonic is collaborating with silicon anode startups like Sila and Nexeon to develop next-generation high-energy-density batteries. In 2023, Panasonic signed a purchase agreement with Nexeon, a UK-based company that produces a range of silicon anode materials with different morphologies and performance characteristics. This partnership will enable Panasonic to produce higher-energy-density EV batteries in the US, with commercial production expected to begin in 2025.
(iv) CATL: The world’s largest battery manufacturer, CATL, is also making significant investments in silicon anode technology. The company’s Qilin battery, which boasts an impressive energy density and fast-charging capability, is rumored to incorporate a silicon-based anode. CATL’s R&D efforts are focused on developing low-cost, high-performance silicon–carbon composite materials. They are also exploring innovative electrolyte formulations and cell designs to improve the overall performance and safety of their silicon-anode batteries. CATL’s vast manufacturing scale and expertise give them a significant advantage in bringing silicon anode technology to the mass market at a competitive cost.

6. Conclusions and Future Perspectives

6.1. Conclusions

This review has systematically examined the landscape of hybrid material systems designed to mitigate the severe volume expansion of silicon anodes, a critical bottleneck hindering their widespread adoption in next-generation lithium-ion batteries. While the theoretical capacity of silicon is an order of magnitude greater than graphite, its practical application is plagued by mechanical degradation, unstable solid electrolyte interphase (SEI) formation, and rapid capacity fade. Our analysis of the literature confirms that hybrid strategies are not merely incremental improvements but a vital and effective approach to creating structurally resilient and electrochemically stable silicon anodes.
Among the various strategies explored, those integrating carbonaceous materials have demonstrated the most significant and consistent success. Specifically, hierarchical structures that provide multi-level protection have proven most effective. For instance, yolk–shell Si@void@C structures and three-dimensional porous graphene frameworks encapsulating silicon nanoparticles have shown exceptional promise. These architectures physically buffer the ~300% volume change by incorporating internal voids, while the external carbon shell provides electrical conductivity and a stable surface for SEI formation. Similarly, sandwich-like structures, such as Si/graphene composites, effectively constrain silicon particles and maintain electrical contact throughout the electrode during cycling. These approaches have led to anodes that can maintain high capacity over hundreds of cycles, a significant improvement over bare silicon.
Despite this progress, critical challenges persist. The initial Coulombic efficiency (ICE) remains a primary obstacle, with much of the initial lithium inventory being consumed in the formation of the SEI layer on the high-surface-area nanostructured silicon. While strategies like pre-lithiation of SiOx@C composites have achieved an impressive ICE of 93.4%, these methods add complexity and cost to the manufacturing process. Furthermore, ensuring the long-term stability of the SEI over thousands of cycles, especially in the presence of continuous volume changes, has not been fully resolved. Finally, the scalability of the most promising synthesis methods remains a major concern. Complex, multi-step processes like chemical vapor deposition (CVD) or atomic layer deposition (ALD), while effective at the lab scale, face significant hurdles in terms of cost, throughput, and industrial feasibility. The development of simpler, more scalable methods, such as the mechanochemical synthesis of Si/Cu3Si composites or the etching of metallurgical-grade Si-Al alloys, represents a crucial and promising direction, but these methods often require further optimization to match the performance of more intricate nanostructures.

6.2. Future Perspectives

Building on the current understanding, future research must pivot from demonstrating proof-of-concept to addressing the practical barriers of commercialization. The path forward requires a multi-faceted approach focused on material design, process engineering, and fundamental understanding.
First, the optimization of hybrid structures must continue, with a focus on balancing electrochemical performance with material and process complexity. Future designs should aim to minimize the content of inactive materials (like carbon shells or metal oxides) without compromising structural integrity. For example, creating covalently bonded Si–carbon interfaces or utilizing multifunctional binders that also participate in charge transport could reduce the reliance on bulky conductive additives and coatings. Computational modeling and machine learning will be instrumental in accelerating the discovery of novel material combinations and predicting the long-term performance of complex architectures, reducing the need for extensive trial-and-error experimentation.
Second, a concerted effort is needed to develop scalable and cost-effective manufacturing processes. While lab-scale synthesis has produced impressive results, the focus must shift to methods compatible with existing battery production lines. Wet-ball milling of industrial-grade silicon with carbon precursors, followed by thermal treatment, is a promising route. Further research into single-step synthesis techniques, such as the direct carbonization of silicon–polymer composites, could dramatically lower production costs. Additionally, developing robust quality control metrics to ensure batch-to-batch consistency of these hybrid materials at an industrial scale is a critical, yet often overlooked, research area.
Third, resolving the interfacial challenges of low ICE and SEI instability is non-negotiable for practical application. Research into advanced electrolyte formulations, including additives that promote the formation of a thin, flexible, and LiF-rich SEI, is paramount. Combining these electrolyte solutions with surface-engineered silicon, such as through ultra-thin ALD coatings of Al2O3 or TiO2, could create a synergistic effect that dramatically enhances cycle life and efficiency. Furthermore, developing practical and scalable pre-lithiation techniques that can be seamlessly integrated into the electrode manufacturing workflow is essential to compensate for the initial lithium loss.
Finally, the research community must move towards evaluating these materials under commercially relevant conditions. This includes testing in full-cell configurations (not just half-cells) with high mass loadings, lean electrolyte, and under a wide range of temperatures and C-rates. Only by validating performance under these stringent conditions can the true potential of hybrid silicon anodes be realized, paving the way for their integration into the next generation of high-energy-density batteries that will power our future.

Author Contributions

Conceptualization, T.K.; methodology, T.K.; writing—original draft preparation, T.K.; writing—review and editing, J.J.; visualization, T.K. and J.J.; supervision, T.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade, Industry & Energy (MOTIE) of the Republic of Korea (No. RS-2024-00509401). This work was partly supported by Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korea government (MOTIE) (No. 20224000000070, Human Resource Training for Smart Energy New Industry Cluster).

Data Availability Statement

Data are gathered from published research articles which are available in public domain.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
LIBLithium-ion battery
EVElectric vehicle
ESSEnergy storage systems
SEISolid electrolyte interphase
CNTsCarbon nanotubes
CNFsCarbon nanofibers
MOxMetal oxides
MXeneMn+1XnTx
(M: early transition metal, X: carbon or nitrogen, Tx: surface functional groups, n = 1–4)
h-BNHexagonal boron nitride
TMDTransition metal dichalcogenide
CPConductive polymer
PANIPolyaniline
PPyPolypyrrole
PEDOT:PPSPoly(3,4-ethylenedioxythiophene):polystyrene sulfonate
CVDChemical vapor deposition
PVDPhysical vapor deposition
ALDAtomic layer deposition
CECoulombic efficiency
ICEInitial Coulombic efficiency

References

  1. Yi, X.; Qi, G.; Liu, X.; Depcik, C.; Liu, L. Challenges and Strategies toward Anode Materials with Different Lithium Storage Mechanisms for Rechargeable Lithium Batteries. J. Energy Storage 2024, 95, 112480. [Google Scholar] [CrossRef]
  2. Wang, K.; Hua, W.; Huang, X.; Stenzel, D.; Wang, J.; Ding, Z.; Cui, Y.; Wang, Q.; Ehrenberg, H.; Breitung, B.; et al. Synergy of Cations in High Entropy Oxide Lithium Ion Battery Anode. Nat. Commun. 2023, 14, 1487. [Google Scholar] [CrossRef]
  3. Shao, G.; Hanaor, D.A.H.; Wang, J.; Kober, D.; Li, S.; Wang, X.; Shen, X.; Bekheet, M.F.; Gurlo, A. Polymer-Derived SiOC Integrated with a Graphene Aerogel as a Highly Stable Li-Ion Battery Anode. ACS Appl. Mater. Interfaces 2020, 12, 46045–46056. [Google Scholar] [CrossRef]
  4. Haro, M.; Kumar, P.; Zhao, J.; Koutsogiannis, P.; Porkovich, A.J.; Ziadi, Z.; Bouloumis, T.; Singh, V.; Juarez-Perez, E.J.; Toulkeridou, E.; et al. Nano-Vault Architecture Mitigates Stress in Silicon-Based Anodes for Lithium-Ion Batteries. Commun. Mater. 2021, 2, 16. [Google Scholar] [CrossRef]
  5. Riedel, O.; Düttmann, A.; Dühnen, S.; Kolny-Olesiak, J.; Gutsche, C.; Parisi, J.; Winter, M.; Knipper, M.; Placke, T. Surface-Modified Tin Nanoparticles and Their Electrochemical Performance in Lithium Ion Battery Cells. ACS Appl. Nano Mater. 2019, 2, 3577–3589. [Google Scholar] [CrossRef]
  6. Han, D.; Wang, Y.; Lv, W.; Wan, B.; Li, Y.; Wang, J.; Sun, S.; Yan, W.; Wang, J. Micro Lithium Batteries toward the Next-generation Smart Microsystems: A Review. Energy Storage Mater. 2025, 81, 104522. [Google Scholar] [CrossRef]
  7. Refino, A.D.; Eldona, C.; Hernandha, R.F.H.; Adhitama, E.; Sumboja, A.; Peiner, E.; Wasisto, H.S. Advances in 3D Silicon-based Lithium-ion Microbatteries. Commun. Mater. 2024, 5, 22. [Google Scholar] [CrossRef]
  8. Refino, A.D.; Yulianto, N.; Priyono, S.; Manawan, M.; Hernandha, R.F.H.; Adhitama, E.; Kottmeier, J.; Dietzel, A.; Peiner, E.; Wasisto, H.S. Carbon-coating Effect on the Performance of Photolithographically-structured Si Nanowires for Lithium-ion Microbattery Anodes. Commun. Mater. 2025, 6, 32. [Google Scholar] [CrossRef]
  9. Cui, Q.; Zhong, Y.; Pan, L.; Zhang, H.; Yang, Y.; Liu, D.; Teng, F.; Bando, Y.; Yao, J.; Wang, X. Recent Advances in Designing High-Capacity Anode Nanomaterials for Li-Ion Batteries and Their Atomic-Scale Storage Mechanism Studies. Adv. Sci. 2018, 5, 1700902. [Google Scholar] [CrossRef] [PubMed]
  10. Je, M.; Han, D.-Y.; Ryu, J.; Park, S. Constructing Pure Si Anodes for Advanced Lithium Batteries. Acc. Chem. Res. 2023, 56, 2213–2224. [Google Scholar] [CrossRef] [PubMed]
  11. Ma, Y.; Guo, P.; Liu, M.; Cheng, P.; Zhang, T.; Liu, J.; Liu, D.; He, D. To Achieve Controlled Specific Capacities of Silicon-Based Anodes for High-Performance Lithium-Ion Batteries. J. Alloys Compd. 2022, 905, 164189. [Google Scholar] [CrossRef]
  12. Sun, L.; Liu, Y.; Wang, L.; Chen, Z.; Jin, Z. Stabilizing Porous Micro-sized Silicon Anodes via Construction of Tough Composite Interface Networks for High-energy-density Lithium-ion Batteries. Nano Res. 2024, 17, 9737–9745. [Google Scholar] [CrossRef]
  13. Li, H.; Yamaguchi, T.; Matsumoto, S.; Hoshikawa, H.; Kumagai, T.; Okamoto, N.L.; Ichitsubo, T. Circumventing Huge Volume Strain in Alloy Anodes of Lithium Batteries. Nat. Commun. 2020, 11, 1584. [Google Scholar] [CrossRef] [PubMed]
  14. Shen, H.; Wang, Q.; Chen, Z.; Rong, C.; Chao, D. Application and Development of Silicon Anode Binders for Lithium-Ion Batteries. Materials 2023, 16, 4266. [Google Scholar] [CrossRef]
  15. Chen, C.-Y.; Sano, T.; Tsuda, T.; Ui, K.; Oshima, Y.; Yamagata, M.; Ishikawa, M.; Haruta, M.; Doi, T.; Inaba, M.; et al. In Situ Scanning Electron Microscopy of Silicon Anode Reactions in Lithium-Ion Batteries during Charge/Discharge Processes. Sci. Rep. 2016, 6, 36153. [Google Scholar] [CrossRef]
  16. Gross, S.J.; Hsieh, M.-T.; Mumm, D.R.; Valdevit, L.; Mohraz, A. Alleviating Expansion-Induced Mechanical Degradation in Lithium-Ion Battery Silicon Anodes via Morphological Design. Extreme Mech. Lett. 2022, 54, 101746. [Google Scholar] [CrossRef]
  17. Müller, S.; Pietsch, P.; Brandt, B.-E.; Baade, P.; De Andrade, V.; De Carlo, F.; Wood, V. Quantification and Modeling of Mechanical Degradation in Lithium-Ion Batteries Based on Nanoscale Imaging. Nat. Commun. 2018, 9, 2340. [Google Scholar] [CrossRef]
  18. Wang, A.; Kadam, S.; Li, H.; Shi, S.; Qi, Y. Review on Modeling of the Anode Solid Electrolyte Interphase (SEI) for Lithium-Ion Batteries. npj Comput. Mater. 2018, 4, 15. [Google Scholar] [CrossRef]
  19. Ezzedine, M.; Jardali, F.; Florea, I.; Zamfir, M.-R.; Cojocaru, C.-S. Nanostructuring Strategies for Silicon-Based Anodes in Lithium-Ion Batteries: Tuning Areal Silicon Loading, SEI Formation/Irreversible Capacity Loss, Rate Capability Retention and Electrode Durability. Batteries Supercaps 2023, 6, e202200451. [Google Scholar] [CrossRef]
  20. Sun, L.; Jiang, X.; Jin, Z. Interfacial Engineering of Porous SiOx@C Composite Anodes toward High-performance Lithium-ion Batteries. Chem. Eng. J. 2023, 474, 145960. [Google Scholar] [CrossRef]
  21. Seydibeyoğlu, M.Ö.; Dogru, A.; Wang, J.; Rencheck, M.; Han, Y.; Wang, L.; Seydibeyoğlu, E.A.; Zhao, X.; Ong, K.; Shatkin, J.A.; et al. Review on Hybrid Reinforced Polymer Matrix Composites with Nanocellulose, Nanomaterials, and Other Fibers. Polymers 2023, 15, 984. [Google Scholar] [CrossRef] [PubMed]
  22. Roviello, G.; Occhicone, A.; De Gregorio, E.; Ricciotti, L.; Cioffi, R.; Ferone, C.; Tarallo, O. Geopolymer-Based Composite and Hybrid Materials: The Synergistic Interaction between Components. Sustain. Mater. Technol. 2025, 44, e01404. [Google Scholar] [CrossRef]
  23. Liang, Y.; Li, Y.; Wang, H.; Dai, H. Strongly Coupled Inorganic/Nanocarbon Hybrid Materials for Advanced Electrocatalysis. J. Am. Chem. Soc. 2013, 135, 2013–2036. [Google Scholar] [CrossRef]
  24. Hou, S.-C.; Chen, T.-Y.; Wu, Y.-H.; Chen, H.-Y.; Lin, X.-D.; Chen, Y.-Q.; Huang, J.-L.; Chang, C.-C. Mechanochemical Synthesis of Si/Cu3Si-Based Composite as Negative Electrode Materials for Lithium Ion Battery. Sci. Rep. 2018, 8, 12695. [Google Scholar] [CrossRef]
  25. Cheng, Y.; Guo, Z.; Zheng, C.; Zhang, L.; Wang, S.; Du, H. Revisiting the Core Problem Impeding the Commercialization of Silicon-Based Lithium-Ion Batteries. Energy Mater. Devices 2025, 3, 9370055. [Google Scholar] [CrossRef]
  26. Zhang, Z.; Ma, R.; Yang, J.G.; Wang, J.; Peng, Y. A Review on Mechanical, Electrical, Chemical, and Electrochemical Properties of Coating Materials for Silicon Anodes in Lithium-Ion Batteries. Small 2025, e06400. [Google Scholar] [CrossRef]
  27. Kong, X.; Xi, Z.; Wang, L.; Zhou, Y.; Liu, Y.; Wang, L.; Li, S.; Chen, X.; Wan, Z. Recent Progress in Silicon-Based Materials for Performance-Enhanced Lithium-Ion Batteries. Molecules 2023, 28, 2079. [Google Scholar] [CrossRef]
  28. Kang, W.; Zhang, Q.; Jia, Y.; Liu, X.; Jiang, N.; Zhao, Y.; Wu, C.; Guan, L. Enhancing the Cycling Stability of Commercial Silicon Nanoparticles by Carbon Coating and Thin Layered Single-Walled Carbon Nanotube Webbing. J. Power Sources 2024, 602, 234338. [Google Scholar] [CrossRef]
  29. Yang, J.; Wang, Y.-X.; Chou, S.-L.; Zhang, R.; Xu, Y.; Fan, J.; Zhang, W.-X.; Liu, H.K.; Zhao, D.; Dou, S.X. Yolk-Shell Silicon-Mesoporous Carbon Anode with Compact Solid Electrolyte Interphase Film for Superior Lithium-Ion Batteries. Nano Energy 2015, 18, 133–142. [Google Scholar] [CrossRef]
  30. Jin, B.; Liao, L.; Shen, X.; Mei, Z.; Du, Q.; Liang, L.; Lei, B.; Du, J. Advancement in Research on Silicon/Carbon Composite Anode Materials for Lithium-Ion Batteries. Metals 2025, 15, 386. [Google Scholar] [CrossRef]
  31. Xu, J.; Yin, Q.; Li, X.; Tan, X.; Liu, Q.; Lu, X.; Cao, B.; Yuan, X.; Li, Y.; Shen, L.; et al. Spheres of Graphene and Carbon Nanotubes Embedding Silicon as Mechanically Resilient Anodes for Lithium-Ion Batteries. Nano Lett. 2022, 22, 3054–3061. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, Z.; Wu, Y.; Mo, Z.; Lei, X.; Xie, X.; Xue, X.; Qin, H.; Jiang, H. Research Progress of Silicon-Based Anode Materials for Lithium-Ion Batteries. RSC Adv. 2025, 15, 10731–10753. [Google Scholar] [CrossRef] [PubMed]
  33. Vorauer, T.; Schöggl, J.; Sanadhya, S.G.; Poluektov, M.; Widanage, W.D.; Figiel, L.; Schädler, S.; Tordoff, B.; Fuchsbichler, B.; Koller, S.; et al. Impact of Solid-Electrolyte Interphase Reformation on Capacity Loss in Silicon-Based Lithium-Ion Batteries. Commun. Mater. 2023, 4, 44. [Google Scholar] [CrossRef]
  34. Li, J.; Fan, S.; Xiu, H.; Wu, H.; Huang, S.; Wang, S.; Yin, D.; Deng, Z.; Xiong, C. TiO2-Coated Silicon Nanoparticle Core–Shell Structure for High-Capacity Lithium-Ion Battery Anode Materials. Nanomaterials 2023, 13, 1144. [Google Scholar] [CrossRef]
  35. Adenusi, H.; Chass, G.A.; Passerini, S.; Tian, K.V.; Chen, G. Lithium Batteries and the Solid Electrolyte Interphase (SEI)—Progress and Outlook. Adv. Energy Mater. 2023, 13, 2203307. [Google Scholar] [CrossRef]
  36. Lin, Y.-X.; Liu, Z.; Leung, K.; Chen, L.-Q.; Lu, P.; Qi, Y. Connecting the Irreversible Capacity Loss in Li-Ion Batteries with the Electronic Insulating Properties of Solid Electrolyte Interphase (SEI) Components. J. Power Sources 2016, 309, 221–230. [Google Scholar] [CrossRef]
  37. Sun, L.; Wang, L.; Wang, T.; Liu, Y.; Qiao, Y.; Lu, X.; Qi, M.; Jin, Z. Fluorinated MXene-engineered LiF-rich Solid Electrolyte Interphase and Hierarchical Confinement Strategy Enabling High Performance Micro-sized Silicon Anodes. Nano Res. 2025. [Google Scholar] [CrossRef]
  38. Toki, G.F.I.; Hossain, M.K.; Rehman, W.U.; Manj, R.Z.A.; Wang, L.; Yang, J. Recent Progress and Challenges in Silicon-Based Anode Materials for Lithium-Ion Batteries. Ind. Chem. Mater. 2024, 2, 226–269. [Google Scholar] [CrossRef]
  39. Sun, L.; Wang, L.; Liu, Y.; Wang, H.; Jin, Z. Synergistic Engineering of Micron-sized Porous Silicon Anodes via Ge Doping and Liquid Metal Alloy Modification for High-energy-density Lithium-ion Batteries. J. Mater. Chem. A 2025, 13, 14346–14352. [Google Scholar] [CrossRef]
  40. Khan, M.; Yan, S.; Ali, M.; Mahmood, F.; Zheng, Y.; Li, G.; Liu, J.; Song, X.; Wang, Y. Innovative Solutions for High Performance Silicon Anodes in Lithium Ion Batteries: Overcoming Challenges and Real World Applications. Nano-Micro Lett. 2024, 16, 179. [Google Scholar] [CrossRef]
  41. Zhou, X.; Wan, L.-J.; Guo, Y.-G. Electrospun Silicon Nanoparticle/Porous Carbon Hybrid Nanofibers for Lithium-Ion Batteries. Small 2013, 9, 2684–2688. [Google Scholar] [CrossRef]
  42. Zhang, M.; Li, J.; Sun, C.; Wang, Z.; Li, Y.; Zhang, D. Durable Flexible Dual-layer and Free-standing Silicon/Carbon Composite Anode for Lithium-ion Batteries. J. Alloys Compd. 2023, 932, 167687. [Google Scholar] [CrossRef]
  43. Li, P.; Miao, C.; Yi, D.; Wei, Y.; Chen, T.; Wu, W. Pomegranate like Silicon–carbon Composites Prepared from Lignin-derived Phenolic Resins as Anode Materials for Lithium-ion Batteries. New J. Chem. 2023, 47, 16855–16863. [Google Scholar] [CrossRef]
  44. Ma, L.; Fu, X.; Zhao, F.; Yu, L.; Su, W.; Wei, L.; Tang, G.; Wang, Y.; Wu, F.; Guo, X. Microsized Silicon/Carbon Composite Anodes through In Situ Polymerization of Phenolic Resin onto Silicon Microparticles for High-Performance Lithium-Ion Batteries. ACS Appl. Energy Mater. 2023, 6, 4989–4999. [Google Scholar] [CrossRef]
  45. Xie, J.; Tong, L.; Su, L.; Xu, Y.; Wang, L.; Wang, Y. Core-shell Yolk-shell Si@C@Void@C Nanohybrids as Advanced Lithium Ion Battery Anodes with Good Electronic Conductivity and Corrosion Resistance. J. Power. Sources 2017, 342, 529–536. [Google Scholar] [CrossRef]
  46. Xu, Z.-L.; Gang, Y.; Garakani, M.A.; Abouali, S.; Huang, J.-Q.; Kim, J.-K. Carbon-coated Mesoporous Silicon Microsphere Anodes with Greatly Reduced Volume Expansion. J. Mater. Chem. A 2016, 4, 6098–6106. [Google Scholar] [CrossRef]
  47. Wang, H.; Xie, J.; Zhang, S.; Cao, G.; Zhao, X. Scalable Preparation of Silicon@Graphite/Carbon Microspheres as High-performance Lithium-ion Battery Anode Materials. RSC Adv. 2016, 6, 69882–69888. [Google Scholar] [CrossRef]
  48. Su, H.; Li, X.; Liu, C.; Shang, Y.; Liu, H. Scalable Synthesis of Micrometer-sized Porous Silicon/carbon Composites for High-stability Lithium-ion Battery Anodes. Chem. Eng. J. 2023, 451, 138394. [Google Scholar] [CrossRef]
  49. Chen, Y.; Du, N.; Zhang, H.; Yang, D. Porous Si@C Coaxial Nanotubes: Layer-by-layer Assembly on ZnO Nanorod Templates and Application to Lithium-ion Batteries. CrystEngComm 2017, 19, 1220–1229. [Google Scholar] [CrossRef]
  50. Wu, J.; Qin, X.; Zhang, H.; He, Y.-B.; Li, B.; Ke, L.; Lv, W.; Du, H.; Yang, Q.-H.; Kang, F. Multilayered Silicon Embedded Porous Carbon/Graphene Hybrid Film as a High Performance Anode. Carbon 2015, 84, 434–443. [Google Scholar] [CrossRef]
  51. Liu, X.; Zhang, J.; Si, W.; Xi, L.; Eichler, B.; Yan, C.; Schmidt, O.G. Sandwich Nanoarchitecture of Si/Reduced Graphene Oxide Bilayer Nanomembranes for Li-Ion Batteries with Long Cycle Life. ACS Nano 2015, 9, 1198–1205. [Google Scholar] [CrossRef]
  52. Sheem, J.M.; Koo, J.K.; Ha, C.; Kim, Y.M.; Kim, Y.U.; Nah, J.H.; Kim, Y.-J. Graphene-Coated Si/C Composites for High-Density Electrodes: Mitigating Silicon Degradation and Enhancing Cycle Life in Lithium-Ion Batteries. Appl. Surf. Sci. Adv. 2025, 26, 100715. [Google Scholar] [CrossRef]
  53. Guo, X.; Yang, Z.; Wang, W.; Zhang, Y.; Yu, N.; Lu, C. Silicon/Carbon Nanotubes Anode for Lithium-Ion Batteries: Synthesis, Interface and Electrochemical Performance. Surf. Interfaces 2024, 48, 104223. [Google Scholar] [CrossRef]
  54. Pei, Y.; Wang, Y.; Chang, A.-Y.; Liao, Y.; Zhang, S.; Wen, X.; Wang, S. Nanofiber-in-Microfiber Carbon/Silicon Composite Anode with High Silicon Content for Lithium-Ion Batteries. Carbon 2023, 203, 436–444. [Google Scholar] [CrossRef]
  55. Zhang, Y.Q.; Xia, X.H.; Wang, X.L.; Mai, Y.J.; Shi, S.J.; Tang, Y.Y.; Li, L.; Tu, J.P. Silicon/Graphene-Sheet Hybrid Film as Anode for Lithium Ion Batteries. Electrochem. Commun. 2012, 23, 17–20. [Google Scholar] [CrossRef]
  56. Feng, Z.; Huang, C.; Fu, A.; Chen, L.; Pei, F.; He, Y.; Fang, X.; Qu, B.; Chen, X.; Ng, A.M.C.; et al. A Three-Dimensional Network of Graphene/Silicon/Graphene Sandwich Sheets as Anode for Li-Ion Battery. Thin Solid Films 2020, 693, 137702. [Google Scholar] [CrossRef]
  57. Cong, R.; Choi, J.-Y.; Song, J.-B.; Jo, M.; Lee, H.; Lee, C.-S. Characteristics and Electrochemical Performances of Silicon/Carbon Nanofiber/Graphene Composite Films as Anode Materials for Binder-Free Lithium-Ion Batteries. Sci. Rep. 2021, 11, 139. [Google Scholar] [CrossRef]
  58. Güneş, F. A Direct Synthesis of Si-Nanowires on 3D Porous Graphene as a High Performance Anode Material for Li-Ion Batteries. RSC Adv. 2016, 6, 1678–1685. [Google Scholar] [CrossRef]
  59. Gohier, A.; Laik, B.; Kim, K.H.; Maurice, J.L.; Pereira-Ramos, J.P.; Cojocaru, C.S.; Van, P.T. High-Rate Capability Silicon Decorated Vertically Aligned Carbon Nanotubes for Li-Ion Batteries. Adv. Mater. 2012, 24, 2592–2597. [Google Scholar] [CrossRef]
  60. Liu, W.; Su, S.; Wang, Y.; Wang, H.; Wang, F.; Wang, G.; Qu, M.; Peng, G.; Xie, Z. Constructing a Stable Conductive Network for High-Performance Silicon-Based Anode in Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2024, 16, 10703–10713. [Google Scholar] [CrossRef]
  61. Zhang, M.; Bai, N.; Lin, W.; Wang, H.; Li, J.; Ma, L.; Wang, X.; Zhang, D.; Cao, Z. Carbon-Encapsulated Silicon Ordered Nanofiber Membranes as High-Performance Anode Material for Lithium-Ion Batteries. J. Alloys Compd. 2025, 1010, 177012. [Google Scholar] [CrossRef]
  62. Zhang, D.; Wang, Y.; Zhao, H.; Cai, C.; Zhang, Z.; Hu, W.; Dong, H.; Ding, S. Embedding Silicon Nanoparticle in Porous Carbon Fiber for Highly Stable Lithium-Ion Battery Anode. Mater. Lett. 2024, 361, 136015. [Google Scholar] [CrossRef]
  63. She, Z.; Gad, M.; Ma, Z.; Li, Y.; Pope, M.A. Enhanced Cycle Stability of Crumpled Graphene-Encapsulated Silicon Anodes via Polydopamine Sealing. ACS Omega 2021, 6, 12293–12305. [Google Scholar] [CrossRef]
  64. Cao, Y.; Su, M.; Zhang, X.; Lin, Q.; Bi, T. A 3D Network-Based Hierarchical Porous Carbon/Si@C Composite as Anode Material for Lithium-Ion Battery with Desirable Performance. J. Energy Storage 2024, 98, 112988. [Google Scholar] [CrossRef]
  65. Kim, S.K.; Chang, H.K.; Kim, C.M.; Yoo, H.D.; Kim, H.S.; Jang, H.D. Fabrication of Ternary Silicon–Carbon Nanotubes–Graphene Composites by Co-Assembly in Evaporating Droplets for Enhanced Electrochemical Energy Storage. J. Alloys Compd. 2018, 751, 43–48. [Google Scholar] [CrossRef]
  66. Min, K.S.; Kim, K.T.; An, H.L.; Go, Y.H.; Lee, Y.E.; Lim, D.W.; Baeck, S.H. Yolk-Shell-Structured SiO2@N,P Co-Doped Carbon Spheres as Highly Stable Anode Materials for Lithium Ion Batteries. J. Power Sources 2022, 543, 231849. [Google Scholar] [CrossRef]
  67. Pappas, G.S.; Ferrari, S.; Huang, X.; Bhagat, R.; Haddleton, D.M.; Wan, C. Heteroatom Doped-Carbon Nanospheres as Anodes in Lithium Ion Batteries. Materials 2016, 9, 35. [Google Scholar] [CrossRef]
  68. Schwan, J.; Nava, G.; Mangolini, L. Critical Barriers to the Large Scale Commercialization of Silicon-containing Batteries. Nanoscale Adv. 2020, 2, 4368–4389. [Google Scholar] [CrossRef]
  69. Kang, H.E.; Ko, J.; Song, S.G.; Yoon, Y.S. Recent Progress in Utilizing Carbon Nanotubes and Graphene to Relieve Volume Expansion and Increase Electrical Conductivity of Si-based Composite Anodes for Lithium-ion Batteries. Carbon 2024, 219, 118800. [Google Scholar] [CrossRef]
  70. Chen, Y.; Yan, Y.; Liu, X.; Zhao, Y.; Wu, X.; Zhou, J.; Wang, Z. Porous Si/Fe2O3 Dual Network Anode for Lithium–Ion Battery Application. Nanomaterials 2020, 10, 2331. [Google Scholar] [CrossRef] [PubMed]
  71. Wu, D.; Shen, R.; Yang, R.; Ji, W.; Jiang, M.; Ding, W.; Peng, L. Mixed Molybdenum Oxides with Superior Performances as an Advanced Anode Material for Lithium-Ion Batteries. Sci. Rep. 2017, 7, 44697. [Google Scholar] [CrossRef]
  72. Zhu, J.; Wang, H.; Lin, R. Improving the Electrochemical Performance of Silicon Materials by SnO2 through Structural Design and Conductivity. Appl. Surf. Sci. 2022, 581, 152230. [Google Scholar] [CrossRef]
  73. Lin, C.W.; Hung, C.L.; Venkateswarlu, M.; Hwang, B.J. Influence of TiO2 Nano-Particles on the Transport Properties of Composite Polymer Electrolyte for Lithium-Ion Batteries. J. Power Sources 2005, 146, 397–401. [Google Scholar] [CrossRef]
  74. Yang, J.; Wang, Y.; Li, W.; Wang, L.; Fan, Y.; Jiang, W.; Luo, W.; Wang, Y.; Kong, B.; Selomulya, C.; et al. Amorphous TiO2 Shells: A Vital Elastic Buffering Layer on Silicon Nanoparticles for High-Performance and Safe Lithium Storage. Adv. Mater. 2017, 29, 1700523. [Google Scholar] [CrossRef]
  75. Li, J.; Wang, Y.; Huang, Z.; Huang, K.; Qi, X.; Zhong, J. Synthesis of Si/TiO2 Core–Shell Nanoparticles as Anode Material for High Performance Lithium Ion Batteries. J. Mater. Sci. Mater. Electron. 2016, 27, 12813–12819. [Google Scholar] [CrossRef]
  76. Li, B.; Zhao, W.; Yang, Z.; Zhang, C.; Dang, F.; Liu, Y.; Jin, F.; Chen, X. A Carbon-Doped Anatase TiO2-Based Flexible Silicon Anode with High-Performance and Stability for Flexible Lithium-Ion Battery. J. Power Sources 2020, 466, 228339. [Google Scholar] [CrossRef]
  77. Lotfabad, E.M.; Kalisvaart, P.; Cui, K.; Kohandehghan, A.; Kupsta, M.; Olsen, B.; Mitlin, D. ALD TiO2 Coated Silicon Nanowires for Lithium Ion Battery Anodes with Enhanced Cycling Stability and Coulombic Efficiency. Phys. Chem. Chem. Phys. 2013, 15, 13646–13657. [Google Scholar] [CrossRef] [PubMed]
  78. Wang, C.; Sireesha, P.; Shang, J.; McCloy, J.S.; Liu, J.; Zhong, W.-H. Converting Commercial Fe2O3 to Effective Anode Material Using Glucose as “Etching” Agent. Ceram. Int. 2023, 49, 32652–32662. [Google Scholar] [CrossRef]
  79. Zhou, M.; Gordin, M.L.; Chen, S.; Xu, T.; Song, J.; Lv, D.; Wang, D. Enhanced Performance of SiO/Fe2O3 Composite as an Anode for Rechargeable Li-Ion Batteries. Electrochem. Commun. 2013, 28, 79–82. [Google Scholar] [CrossRef]
  80. Yan, Y.; Chen, Y.; Li, Y.; Wu, X.; Jin, C.; Wang, Z. Synthesis of Si/Fe2O3-Anchored rGO Frameworks as High-Performance Anodes for Li-Ion Batteries. Int. J. Mol. Sci. 2021, 22, 11041. [Google Scholar] [CrossRef]
  81. Wang, Q.; Guo, C.; He, J.; Yang, S.; Liu, Z.; Wang, Q. Fe2O3/C-Modified Si Nanoparticles as Anode Material for High-Performance Lithium-Ion Batteries. J. Alloys Compd. 2019, 795, 284–290. [Google Scholar] [CrossRef]
  82. Wang, B.; Su, D.; Park, J.; Ahn, H.; Wang, G. Graphene-Supported SnO2 Nanoparticles Prepared by a Solvothermal Approach for an Enhanced Electrochemical Performance in Lithium-Ion Batteries. Nanoscale Res. Lett. 2012, 7, 215. [Google Scholar] [CrossRef] [PubMed]
  83. Lei, Y.; Li, S.; Du, M.; Mi, J.; Gao, D.-C.; Hao, L.; Jiang, L.-J.; Luo, M.; Jiang, W.-Q.; Li, F.; et al. Preparation of Double-Shell Si@SnO2@C Nanocomposite as Anode for Lithium-Ion Batteries by Hydrothermal Method. Rare Met. 2023, 42, 2972–2981. [Google Scholar] [CrossRef]
  84. Ma, T.; Yu, X.; Li, H.; Zhang, W.; Cheng, X.; Zhu, W.; Qiu, X. High Volumetric Capacity of Hollow Structured SnO2@Si Nanospheres for Lithium-Ion Batteries. Nano Lett. 2017, 17, 3959–3964. [Google Scholar] [CrossRef]
  85. Li, C.; Chen, Y.; Wei, B.; Ding, K.; Zhang, Y.; Shi, X.; Zhou, J. A Novel Composite Anode Material of Si–SnO2–Graphene Prepared in Air for Lithium Ion Batteries. Int. J. Electrochem. Sci. 2017, 12, 11701–11714. [Google Scholar] [CrossRef]
  86. Han, J.; Kong, D.; Lv, W.; Tang, D.-M.; Han, D.; Zhang, C.; Liu, D.; Xiao, Z.; Zhang, X.; Xiao, J.; et al. Caging Tin Oxide in Three-dimensional Graphene Networks for Superior Volumetric Lithium Storage. Nat. Commun. 2018, 9, 402. [Google Scholar] [CrossRef]
  87. Yuca, N.; Kalafat, I.; Guney, E.; Cetin, B.; Taskin, O.S. Self-Healing Systems in Silicon Anodes for Li-Ion Batteries. Materials 2022, 15, 2392. [Google Scholar] [CrossRef]
  88. Wenelska, K.; Ottmann, A.; Schneider, P.; Thauer, E.; Klingeler, R.; Mijowska, E. Hollow Carbon Sphere/Metal Oxide Nanocomposites Anodes for Lithium-Ion Batteries. Energy 2016, 103, 100–106. [Google Scholar] [CrossRef]
  89. Huertas, Z.C.; Settipani, D.; Flox, C.; Morante, J.R.; Kallio, T.; Biendicho, J.J. High Performance Silicon Electrode Enabled by Titanicone Coating. Sci. Rep. 2022, 12, 137. [Google Scholar] [CrossRef] [PubMed]
  90. Jiang, T.; Yang, H.; Chen, G.Z. Enhanced Performance of Silicon Negative Electrodes Composited with Titanium Carbide Based MXenes for Lithium-Ion Batteries. Nanoenergy Adv. 2022, 2, 165–196. [Google Scholar] [CrossRef]
  91. Kong, F.; He, X.; Liu, Q.; Qi, X.; Sun, D.; Zheng, Y.; Wang, R.; Bai, Y. Enhanced Reversible Li-ion Storage in Si@Ti3C2 MXene Nanocomposite. Electrochem. Commun. 2018, 97, 16–21. [Google Scholar] [CrossRef]
  92. Yang, Q.; Wang, Z.; Xia, Y.; Wu, G.; Chen, C.; Wang, J.; Rao, P.; Dong, A. Facile Electrostatic Assembly of Si@MXene Superstructures for Enhanced Lithium-ion Storage. J. Colloid Interface Sci. 2020, 580, 68–76. [Google Scholar] [CrossRef]
  93. Zhang, C.J.; Park, S.-H.; Seral-Ascaso, A.; Barwich, S.; McEvoy, N.; Boland, C.S.; Coleman, J.N.; Gogotsi, Y.; Nicolosi, V. High Capacity Silicon Anodes Enabled by MXene Viscous Aqueous Ink. Nat. Commun. 2019, 10, 849. [Google Scholar] [CrossRef]
  94. Tian, Y.; An, Y.; Feng, J. Flexible and Freestanding Silicon/MXene Composite Papers for High-Performance Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2019, 11, 10004–10011. [Google Scholar] [CrossRef]
  95. Marriam, I.; Tebyetekerwa, M.; Chen, H.; Chathuranga, H.; Motta, N.; Alarco, J.A.; He, Z.-J.; Zheng, J.-C.; Du, A.; Yan, C. Few-Layer MoS2 Nanosheets with and without Silicon Nanoparticles as Anodes for Lithium-Ion Batteries. J. Mater. Chem. A 2023, 11, 2670–2678. [Google Scholar] [CrossRef]
  96. Gogotsi, Y.; Huang, Q. MXenes: Two-Dimensional Building Blocks for Future Materials and Devices. ACS Nano 2021, 15, 5775–5780. [Google Scholar] [CrossRef] [PubMed]
  97. Zhu, X.; Shen, J.; Chen, X.; Li, Y.; Peng, W.; Zhang, G.; Zhang, F.; Fan, X. Enhanced Cycling Performance of Si–MXene Nanohybrids as Anode for High Performance Lithium Ion Batteries. Chem. Eng. J. 2019, 378, 122212. [Google Scholar] [CrossRef]
  98. Kong, S.; Liu, C.; Ren, J.; Wang, T.; Geng, X.; Yuan, Y.; Zhao, C.; Zhao, C.; Yang, L. The Synergistic Effect of Cross-Linked and Electrostatic Self-Assembly Si/MXene Composites Anode for Highly Efficient Lithium-Ion Battery. Coatings 2024, 14, 1210. [Google Scholar] [CrossRef]
  99. Zhou, Y.; Kucheyev, S.O.; Wan, L.F. First-Principles Elucidation of Defect-Mediated Li Transport in Hexagonal Boron Nitride. Phys. Chem. Chem. Phys. 2025, 27, 3997–4003. [Google Scholar] [CrossRef]
  100. Yadav, D.; Jung, J.-H.; Lee, Y.; Kim, T.Y.-S.; Park, E.; Choi, K.-I.; Cha, J.; Song, W.-J.; Choi, J.-H.; Doo, S.; et al. Enhancement of Coulombic Efficiency and Capacity of Li-Ion Batteries Using a Boron Nitride Nanotubes-Dispersed Electrolyte with High Ionic Conductivity. ACS Mater. Lett. 2023, 5, 2648–2655. [Google Scholar] [CrossRef]
  101. Song, W.; Chae, O.B. Surface-Coating Strategies of Si-Negative Electrode Materials in Lithium-Ion Batteries. Batteries 2024, 10, 327. [Google Scholar] [CrossRef]
  102. Mayorga-Martinez, C.C.; Ambrosi, A.; Eng, A.Y.S.; Sofer, Z.; Pumera, M. Transition Metal Dichalcogenides (MoS2, MoSe2, WS2 and WSe2) Exfoliation Technique Has Strong Influence upon Their Capacitance. Electrochem. Commun. 2015, 56, 24–28. [Google Scholar] [CrossRef]
  103. Kawade, U.V.; Ambalkar, A.A.; Panmand, R.P.; Kalubarme, R.S.; Kadam, S.R.; Naik, S.D.; Kulkarni, M.V.; Kale, B.B. Silicon Nanoparticle-Sandwiched Ultrathin MoS2–Graphene Layers as an Anode Material for Li-Ion Batteries. Mater. Chem. Front. 2019, 3, 587–596. [Google Scholar] [CrossRef]
  104. Wang, J.; Malgras, V.; Sugahara, Y.; Yamauchi, Y. Electrochemical Energy Storage Performance of 2D Nanoarchitectured Hybrid Materials. Nat. Commun. 2021, 12, 3563. [Google Scholar] [CrossRef]
  105. He, Y.; Shen, X.; Zhang, Y. Layered 2D Materials in Batteries. ACS Appl. Nano Mater. 2024, 7, 27907–27939. [Google Scholar] [CrossRef]
  106. Ka, N.; Rout, C.S. Conducting Polymers: A Comprehensive Review on Recent Advances in Synthesis, Properties and Applications. RSC Adv. 2021, 11, 5659–5697. [Google Scholar] [CrossRef]
  107. Balqis, F.; Eldona, C.; Laksono, B.T.; Aini, Q.; Hamid, F.H.; Wasisto, H.S.; Sumboja, A. Conductive Polymer Frameworks in Silicon Anodes for Advanced Lithium-Ion Batteries. ACS Appl. Polym. Mater. 2023, 5, 4933–4952. [Google Scholar] [CrossRef]
  108. Wu, H.; Yu, G.; Pan, L.; Liu, N.; McDowell, M.T.; Bao, Z.; Cui, Y. Stable Li-Ion Battery Anodes by In-Situ Polymerization of Conducting Hydrogel to Conformally Coat Silicon Nanoparticles. Nat. Commun. 2013, 4, 1943. [Google Scholar] [CrossRef]
  109. Zhou, J.; Zhou, L.; Yang, L.; Chen, T.; Li, J.; Pan, H.; Yang, Y.; Wang, Z. Carbon Free Silicon/Polyaniline Hybrid Anodes with 3D Conductive Structures for Superior Lithium-Ion Batteries. Chem. Commun. 2020, 56, 2328–2331. [Google Scholar] [CrossRef]
  110. Zhang, S.; Chen, S.; Wang, Y.; Zhang, T.; Yue, H.; Li, T.; Li, W.; Li, H.; Hao, Y.; Gao, Y. Fabrication of Polypyrrole-Coated Silicon Nanoparticle Composite Electrode for Lithium-Ion Battery. Ionics 2024, 30, 7869–7879. [Google Scholar] [CrossRef]
  111. Yue, L.; Wang, S.; Zhao, X.; Zhang, L. Nano-Silicon Composites Using Poly(3,4-Ethylenedioxythiophene):Poly(styrenesulfonate) as Elastic Polymer Matrix and Carbon Source for Lithium-Ion Battery Anode. J. Mater. Chem. 2012, 22, 1094–1099. [Google Scholar] [CrossRef]
  112. Bednarczyk, K.; Matysiak, W.; Tański, T.; Janeczek, H.; Schab-Balcerzak, E.; Libera, M. Effect of Polyaniline Content and Protonating Dopants on Electroconductive Composites. Sci. Rep. 2021, 11, 7487. [Google Scholar] [CrossRef]
  113. Tu, J.; Hu, L.; Wang, W.; Hou, J.; Zhu, H.; Jiao, S. In-Situ Synthesis of Silicon/Polyaniline Core/Shell and Its Electrochemical Performance for Lithium-Ion Batteries. J. Electrochem. Soc. 2013, 160, A1916–A1921. [Google Scholar] [CrossRef]
  114. Mu, G.; Ding, Z.; Mu, D.; Wu, B.; Bi, J.; Zhang, L.; Yang, H.; Wu, H.; Wu, F. Hierarchical Void Structured Si/PANi/C Hybrid Anode Material for High-Performance Lithium-Ion Batteries. Electrochim. Acta 2019, 300, 341–348. [Google Scholar] [CrossRef]
  115. Eldona, C.; Hawari, N.H.; Hamid, F.H.; Dempwolf, W.; Iskandar, F.; Peiner, E.; Wasisto, H.S.; Sumboja, A. A Free-Standing Polyaniline/Silicon Nanowire Forest as the Anode for Lithium-Ion Batteries. Chem. Asian J. 2022, 17, e202200946. [Google Scholar] [CrossRef] [PubMed]
  116. Cai, J.-J.; Zuo, P.-J.; Cheng, X.-Q.; Xu, Y.-H.; Yin, G.-P. Nano-Silicon/Polyaniline Composite for Lithium Storage. Electrochem. Commun. 2010, 12, 1572–1575. [Google Scholar] [CrossRef]
  117. Pang, A.L.; Arsad, A.; Ahmadipour, M. Synthesis and Factor Affecting on the Conductivity of Polypyrrole: A Short Review. Polym. Adv. Technol. 2021, 32, 1428–1454. [Google Scholar] [CrossRef]
  118. Guo, Z.P.; Wang, J.Z.; Liu, H.K.; Dou, S.X. Study of Silicon/Polypyrrole Composite as Anode Materials for Li-Ion Batteries. J. Power Sources 2005, 146, 448–451. [Google Scholar] [CrossRef]
  119. Du, F.-H.; Li, B.; Fu, W.; Xiong, Y.-J.; Wang, K.-X.; Chen, J.-S. Surface Binding of Polypyrrole on Porous Silicon Hollow Nanospheres for Li-Ion Battery Anodes with High Structure Stability. Adv. Mater. 2014, 26, 6145–6150. [Google Scholar] [CrossRef]
  120. Kim, Y.; Yoo, S.; Kim, J.-H. Water-Based Highly Stretchable PEDOT:PSS/Nonionic WPU Transparent Electrode. Polymers 2022, 14, 949. [Google Scholar] [CrossRef] [PubMed]
  121. Li, S.; Huang, J.; Wang, J.; Han, K. Micro-Sized Porous Silicon@PEDOT with High Rate Capacity and Stability for Li-Ion Battery Anode. Mater. Lett. 2021, 293, 129712. [Google Scholar] [CrossRef]
  122. Higgins, T.M.; Park, S.-H.; King, P.J.; Zhang, C.; McEvoy, N.; Berner, N.C.; Daly, D.; Shmeliov, A.; Khan, U.; Duesberg, G.; et al. A Commercial Conducting Polymer as Both Binder and Conductive Additive for Silicon Nanoparticle-Based Lithium-Ion Battery Negative Electrodes. ACS Nano 2016, 10, 3702–3713. [Google Scholar] [CrossRef]
  123. Dhason, M.V.A.; Bhattacharya, I.; Mansour, M.; Lazer, S.J.; Banik, T.; Soyoye, B. Conductive Polymers in Si Anodes for Lithium-Ion Batteries: Advancements, Challenges and Future Aspects. Mater. Today Energy 2025, 51, 101897. [Google Scholar] [CrossRef]
  124. Maji, R.; Salvador, M.A.; Ruini, A.; Magri, R.; Degoli, E. A First-Principles Study of Self-Healing Binders for Next-Generation Si-Based Lithium-Ion Batteries. Mater. Today Chem. 2023, 29, 101474. [Google Scholar] [CrossRef]
  125. Hwa, Y.; Kim, W.-S.; Hong, S.-H.; Sohn, H.-J. High Capacity and Rate Capability of Core–Shell Structured Nano-Si/C Anode for Li-Ion Batteries. Electrochim. Acta 2012, 71, 201–205. [Google Scholar] [CrossRef]
  126. Zeng, Q.; Huang, Q.; Luo, Z.; Liu, A.; Li, H.; Zhong, M.; Zhang, Z. Sandwich Network Structure Silicon/Carbon Anode Material for Lithium-Ion Batteries Based on Bacterial Cellulose. Ionics 2024, 30, 135–144. [Google Scholar] [CrossRef]
  127. Jiang, Y.; Chen, S.; Mu, D.; Wu, B.; Liu, Q.; Zhao, Z.; Wu, F. A Three-Dimensional Network Structure Si/C Anode for Li-Ion Batteries. J. Mater. Sci. 2017, 52, 10950–10958. [Google Scholar] [CrossRef]
  128. Zhou, X.-Y.; Tang, J.-J.; Yang, J.; Xie, J.; Ma, L.-L. Silicon@Carbon Hollow Core–Shell Heterostructures Novel Anode Materials for Lithium Ion Batteries. Electrochim. Acta 2013, 87, 663–668. [Google Scholar] [CrossRef]
  129. Xu, Y.; Yin, G.; Ma, Y.; Zuo, P.; Cheng, X. Nanosized Core/Shell Silicon@Carbon Anode Material for Lithium Ion Batteries with Polyvinylidene Fluoride as Carbon Source. J. Mater. Chem. 2010, 20, 3216–3220. [Google Scholar] [CrossRef]
  130. Hsu, Y.-C.; Hsieh, C.-C.; Liu, W.-R. Synthesis of Double Core–Shell Carbon/Silicon/Graphite Composite Anode Materials for Lithium-Ion Batteries. Surf. Coat. Technol. 2020, 387, 125528. [Google Scholar] [CrossRef]
  131. Qiao, Y.; Hu, Y.; Qian, Z.; Qu, M.; Liu, Z. An Innovative Strategy for Constructing Multicore Yolk-Shell Si/C Anodes for Lithium-Ion Batteries. J. Colloid Interface Sci. 2025, 684, 678–689. [Google Scholar] [CrossRef]
  132. Bi, X.; Tang, T.; Shi, X.; Ge, X.; Wu, W.; Zhang, Z.; Wang, J. One-Step Synthesis of Multi-Core–Void@Shell Structured Silicon Anode for High-Performance Lithium-Ion Batteries. Small 2022, 18, 2200796. [Google Scholar] [CrossRef]
  133. Wang, F.; Wang, B.; Ruan, T.; Gao, T.; Song, R.; Jin, F.; Zhou, Y.; Wang, D.; Liu, H.; Dou, S. Construction of Structure-Tunable Si@Void@C Anode Materials for Lithium-Ion Batteries through Controlling the Growth Kinetics of Resin. ACS Nano 2019, 13, 12219–12229. [Google Scholar] [CrossRef]
  134. Shi, J.; Gao, H.; Hu, G.; Zhang, Q. Interfacial Self-Assembled Si@SiOx@C Microclusters with High Tap Density for High-Performance Li-Ion Batteries. Mater. Today Energy 2022, 29, 101090. [Google Scholar] [CrossRef]
  135. Xiang, J.; Liu, H.; Na, R.; Wang, D.; Shan, Z.; Tian, J. Facile Preparation of Void-Buffered Si@TiO2/C Microspheres for High-Capacity Lithium Ion Battery Anodes. Electrochim. Acta 2020, 337, 135841. [Google Scholar] [CrossRef]
  136. Lee, J.-I.; Ko, Y.; Shin, M.; Song, H.-K.; Choi, N.-S.; Kim, M.G.; Park, S. High-performance Silicon-based Multicomponent Battery Anodes Produced via Synergistic Coupling of Multifunctional Coating Layers. Energy Environ. Sci. 2015, 8, 2075–2084. [Google Scholar] [CrossRef]
  137. An, W.; He, P.; Xiao, C.; Guo, E.; Pang, C.; He, X.; Ren, J.; Yuan, G.; Du, N.; Yang, D. Hierarchical Carbon Shell Compositing Microscale Silicon Skeleton as High-Performance Anodes for Lithium-Ion Batteries. ACS Appl. Energy Mater. 2021, 4, 4976–4985. [Google Scholar] [CrossRef]
  138. Wang, L.; Wang, X.; Wang, T.; Liu, Y.; Sun, L. Mechanically Robust and Interface-Stable pSi@GO–CNTs@C Anode via Multi-Level Confinement for Long-Cycling Lithium-Ion Batteries. ACS Appl. Energy Mater. 2025, 8, 12920–12928. [Google Scholar] [CrossRef]
  139. Wang, X.; Li, Y.; Zhang, B.; Chen, Y.; Liu, H. Enhanced Electrochemical Performance of Si/C Anode by a Facile Surface Modification Strategy. Appl. Surf. Sci. 2017, 425, 614–621. [Google Scholar]
  140. Chen, Y.; Huang, J.; Chen, Z.; Zeng, H.; Wu, Z.; Yang, H.; Chen, L.; Sun, Q.; Qian, W. Electrospun Sandwich-Structured C@Si/C@C as Anode for Advanced Lithium-Ion Batteries. Appl. Surf. Sci. 2025, 699, 163139. [Google Scholar] [CrossRef]
  141. Yang, Z.; Xia, Y.; Ji, J.; Qiu, B.; Zhang, K.; Liu, Z. Superior Cycling Performance of a Sandwich Structure Si/C Anode for Lithium Ion Batteries. RSC Adv. 2016, 6, 12107–12113. [Google Scholar] [CrossRef]
  142. Yang, Y.; Zhang, R.; Zhang, Q.; Feng, L.; Wen, G.; Qin, L.-C.; Wang, D. Using Sandwiched Silicon/Reduced Graphene Oxide Composites as Anodes for Lithium-Ion Batteries. Molecules 2024, 29, 2178. [Google Scholar] [CrossRef] [PubMed]
  143. Ko, K.-Y.; Sung, K.-W.; Ahn, H.-J. Layer-by-Layer-Structured Silicon-Based Electrode Design for Ultrafast Lithium-Ion Batteries. Korean J. Chem. Eng. 2025, 42, 1045–1053. [Google Scholar] [CrossRef]
  144. Chang, J.; Huang, X.; Zhou, G.; Cui, S.; Hallac, P.B.; Jiang, J.; Hurley, P.T.; Chen, J. Multilayered Si Nanoparticle/Reduced Graphene Oxide Hybrid as a High-Performance Lithium-Ion Battery Anode. Adv. Mater. 2014, 26, 758–764. [Google Scholar] [CrossRef]
  145. Zhang, Y.; Tang, W.; Gao, H.; Li, M.; Wan, H.; Kong, X.; Liu, X.; Chen, G.; Chen, Z. Monolithic Layered Silicon Composed of a Crystalline–Amorphous Network for Sustainable Lithium-Ion Battery Anodes. ACS Nano 2024, 18, 15671–15680. [Google Scholar] [CrossRef]
  146. Di, F.; Gu, X.; Chu, Y.; Li, L.; Geng, X.; Sun, C.; Zhou, W.; Zhang, H.; Zhao, H.; Tao, L.; et al. Enhanced Stability and Kinetic Performance of Sandwich Si Anode Constructed by Carbon Nanotube and Silicon Carbide for Lithium-Ion Battery. J. Colloid Interface Sci. 2024, 670, 204–214. [Google Scholar] [CrossRef]
  147. Fang, H.; Liu, Q.; Feng, X.; Yan, J.; Wang, L.; He, L.; Zhang, L.; Wang, G. Carbon-Coated Si Nanoparticles Anchored on Three-Dimensional Carbon Nanotube Matrix for High-Energy Stable Lithium-Ion Batteries. Batteries 2023, 9, 118. [Google Scholar] [CrossRef]
  148. Yang, B.; Liu, F.; Liu, Y.; Dong, J.; Liu, M.; Wang, S.; Zhang, L. Self-Assembled Three-Dimensional Si/Carbon Frameworks as Promising Lithium-Ion Battery Anode. J. Power Sources 2023, 553, 232274. [Google Scholar] [CrossRef]
  149. Zhang, K.; Xia, Y.; Yang, Z.; Fu, R.; Shen, C.; Liu, Z. Structure-Preserved 3D Porous Silicon/Reduced Graphene Oxide Materials as Anodes for Li-Ion Batteries. RSC Adv. 2017, 7, 24305–24311. [Google Scholar] [CrossRef]
  150. Li, X.; Chen, Z.; Li, A.; Yu, Y.; Chen, X.; Song, H. Three-Dimensional Hierarchical Porous Structures Constructed by Two-Stage MXene-Wrapped Si Nanoparticles for Li-Ion Batteries. ACS Appl. Mater. Interfaces 2020, 12, 48718–48728. [Google Scholar] [CrossRef] [PubMed]
  151. Liu, G.; Zhu, X.; Li, X.; Jia, D.; Li, D.; Ma, Z.; Li, J. Flexible Porous Silicon/Carbon Fiber Anode for High−Performance Lithium−Ion Batteries. Materials 2022, 15, 3190. [Google Scholar] [CrossRef]
  152. Chen, W.-T.; Muruganantham, R.; Liu, W.-R. Construction of 3D Porous Graphene Aerogel Wrapped Silicon Composite as Anode Materials for High-Efficient Lithium-Ion Storage. Surf. Coat. Technol. 2022, 434, 128147. [Google Scholar] [CrossRef]
  153. Wang, W.; Favors, Z.; Ionescu, R.; Ye, R.; Hosseini Bay, H.; Ozkan, M.; Ozkan, C.S. Monodisperse Porous Silicon Spheres as Anode Materials for Lithium Ion Batteries. Sci. Rep. 2015, 5, 8781. [Google Scholar] [CrossRef]
  154. Ji, L.; Zhang, X. Electrospun Carbon Nanofibers Containing Silicon Particles as an Energy-Storage Medium. Carbon 2009, 47, 3219–3226. [Google Scholar] [CrossRef]
  155. Li, Q.; Yin, L.; Gao, X. Reduction Chemical Reaction Synthesized Scalable 3D Porous Silicon/Carbon Hybrid Architectures as Anode Materials for Lithium Ion Batteries with Enhanced Electrochemical Performance. RSC Adv. 2015, 5, 35598–35607. [Google Scholar] [CrossRef]
  156. Wu, Y.; Guo, Z.-S. Modeling Li-Ion Concentration Distribution and Stress of Porous Electrode Particles Considering Binder and Direct Particle Contact. J. Energy Storage 2021, 44, 103315. [Google Scholar] [CrossRef]
  157. Zhang, R.; Yu, P.; Li, Z.; Shen, X.; Yu, Y.; Yu, J. Hierarchical Porous Structured Si/C Anode Material for Lithium-Ion Batteries by Dual Encapsulating Layers for Enhanced Lithium-Ion and Electron Transports Rates. Small 2025, 21, 2407276. [Google Scholar] [CrossRef]
  158. Xiao, Q.; Gu, M.; Yang, H.; Li, B.; Zhang, C.; Liu, Y.; Liu, F.; Dai, F.; Yang, L.; Liu, Z.; et al. Inward Lithium-Ion Breathing of Hierarchically Porous Silicon Anodes. Nat. Commun. 2015, 6, 8844. [Google Scholar] [CrossRef] [PubMed]
  159. Chen, S.; Bao, P.; Huang, X.; Sun, B.; Wang, G. Hierarchical 3D Mesoporous Silicon@Graphene Nanoarchitectures for Lithium Ion Batteries with Superior Performance. Nano Res. 2014, 7, 85–94. [Google Scholar] [CrossRef]
  160. Tagliaferri, S.; Gaspard, L.; Au, H.; Mattevi, C.; Titirici, M.-M.; Crespo-Ribadeneyra, M. Nature-Inspired Batteries: From Biomaterials to Biomimetic Design Strategies. Green Chem. 2024, 26, 6944–6958. [Google Scholar] [CrossRef]
  161. Beydaghi, H.; Abouali, S.; Thorat, S.B.; Del Rio Castillo, A.E.; Bellani, S.; Lauciello, S.; Gentiluomo, S.; Pellegrini, V.; Bonaccorso, F. 3D Printed Silicon–Few Layer Graphene Anode for Advanced Li-Ion Batteries. RSC Adv. 2021, 11, 35051–35060. [Google Scholar] [CrossRef]
  162. Suryanarayana, C. Mechanical Alloying and Milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  163. Liang, Y.-Z.; Bhat, A.L.; Su, Y.-S. Green Synthesis of Graphene Flake/Silicon Composite Anode for Lithium-Ion Batteries Using a Ball-Mill-Derived Mechanical Transfer Technique. ACS Appl. Energy Mater. 2024, 7, 10574–10583. [Google Scholar] [CrossRef]
  164. Tiano, A.L.; Koenigsmann, C.; Santulli, A.C.; Wong, S.S. Solution-Based Synthetic Strategies for One-Dimensional Metal-Containing Nanostructures. Chem. Commun. 2010, 46, 8093–8130. [Google Scholar] [CrossRef] [PubMed]
  165. Hwa, Y.; Kim, W.-S.; Yu, B.-C.; Hong, S.-H.; Sohn, H.-J. Enhancement of the Cyclability of a Si Anode through Co3O4 Coating by the Sol–Gel Method. J. Phys. Chem. C 2013, 117, 7013–7017. [Google Scholar] [CrossRef]
  166. Ruttert, M.; Holtstiege, F.; Hüsker, J.; Börner, M.; Winter, M.; Placke, T. Hydrothermal-Derived Carbon as a Stabilizing Matrix for Improved Cycling Performance of Silicon-Based Anodes for Lithium-Ion Full Cells. Beilstein J. Nanotechnol. 2018, 9, 2381–2395. [Google Scholar] [CrossRef]
  167. Chen, Y.; Nie, M.; Lucht, B.L.; Saha, A.; Guduru, P.R.; Bose, A. High Capacity, Stable Silicon/Carbon Anodes for Lithium-Ion Batteries Prepared Using Emulsion-Templated Directed Assembly. ACS Appl. Mater. Interfaces 2014, 6, 4678–4683. [Google Scholar] [CrossRef]
  168. Yang, Y.; Liu, Y.; Jiang, X.; Zhao, L.; Wang, P.; Zhang, Y. Rational Design of SiOx Based Anode Materials for Next Generation Lithium-ion Batteries. Mater. Adv. 2024, 5, 896–919. [Google Scholar] [CrossRef]
  169. Wan, W.; Mai, Y.; Guo, D.; Hou, G.; Dai, X.; Gu, Y.; Li, S.; Wu, F. A Novel Sol-gel Process to Encapsulate Micron Silicon with a Uniformly Ni-doped Graphite Carbon Layer by Coupling for Use in Lithium Ion Batteries. Synth. Met. 2021, 274, 116717. [Google Scholar] [CrossRef]
  170. Zhang, W.; Li, J.; Guan, P.; Lv, C.; Yang, C.; Han, N.; Wang, X.; Song, G.; Peng, Z. One-Pot Sol-Gel Synthesis of Si/C Yolk-Shell Anodes for High Performance Lithium-Ion Batteries. J. Alloys Compd. 2020, 835, 155135. [Google Scholar] [CrossRef]
  171. Shandilya, M.; Rai, R.; Singh, J. Review: Hydrothermal Technology for Smart Materials. Int. Mater. Rev. 2016, 61, 477–509. [Google Scholar] [CrossRef]
  172. Zhang, Y.; Zhao, Y.; Niu, Y.; Ren, J.; Hou, H. Facile Hydrothermal Synthesis of Double Shelled Si@SnO2@C as Advanced Cathode for High-Temperature Lithium Batteries. J. Alloys Compd. 2021, 858, 157661. [Google Scholar] [CrossRef]
  173. Mery, A.; Chenavier, Y.; Marcucci, C.; Benayad, A.; Alper, J.P.; Dubois, L.; Haon, C.; Herlin-Boime, N.; Sadki, S.; Duclairoir, F. Toward the Improvement of Silicon-Based Composite Electrodes via an In-Situ Si@C-Graphene Composite Synthesis for Li-Ion Battery Applications. Materials 2023, 16, 2451. [Google Scholar] [CrossRef]
  174. Qin, J.; Wu, M.; Feng, T.; Chen, C.; Tu, C.; Li, X.; Duan, C.; Xia, D.; Wang, D. High Rate Capability and Long Cycling Life of Graphene-Coated Silicon Composite Anodes for Lithium Ion Batteries. Electrochim. Acta 2017, 256, 259–266. [Google Scholar] [CrossRef]
  175. Malik, M.A.; Wani, M.Y.; Hashim, M.A. Microemulsion Method: A Novel Route to Synthesize Organic and Inorganic Nanomaterials: 1st Nano Update. Arab. J. Chem. 2012, 5, 397–417. [Google Scholar] [CrossRef]
  176. Finnie, K.S.; Bartlett, J.R.; Barbé, C.J.A.; Kong, L. Formation of Silica Nanoparticles in Microemulsions. Langmuir 2007, 23, 3017–3024. [Google Scholar] [CrossRef] [PubMed]
  177. Zhang, T.; Fu, L.; Gao, J.; Yang, L.; Wu, Y.; Wu, H. Core-Shell Si/C Nanocomposite as Anode Material for Lithium Ion Batteries. Pure Appl. Chem. 2006, 78, 1889–1895. [Google Scholar] [CrossRef]
  178. Lajoie, L.; Fabiano-Tixier, A.-S.; Chemat, F. Water as Green Solvent: Methods of Solubilisation and Extraction of Natural Products—Past, Present and Future Solutions. Pharmaceuticals 2022, 15, 1507. [Google Scholar] [CrossRef] [PubMed]
  179. Sun, Y.; Fujisawa, K.; Lin, Z.; Lei, Y.; Mondschein, J.S.; Terrones, M.; Schaak, R.E. Low-Temperature Solution Synthesis of Transition Metal Dichalcogenide Alloys with Tunable Optical Properties. J. Am. Chem. Soc. 2017, 139, 11096–11105. [Google Scholar] [CrossRef]
  180. Xu, Z.; Yu, Y.; Fang, D.; Xu, J.; Liang, J.; Zhou, L. Microwave–Ultrasound Assisted Synthesis of β-FeOOH and Its Catalytic Property in a Photo-Fenton-Like Process. Ultrason. Sonochem. 2015, 27, 287–295. [Google Scholar] [CrossRef]
  181. Barrino, F. Hybrid Organic–Inorganic Materials Prepared by Sol–Gel and Sol–Gel-Coating Method for Biomedical Use: Study and Synthetic Review of Synthesis and Properties. Coatings 2024, 14, 425. [Google Scholar] [CrossRef]
  182. Yu, C.; Li, X.; Ma, T.; Rong, J.; Zhang, R.; Shaffer, J.; An, Y.; Liu, Q.; Wei, B.; Jiang, H. Silicon Thin Films as Anodes for High-Performance Lithium-Ion Batteries with Effective Stress Relaxation. Adv. Energy Mater. 2012, 2, 68–73. [Google Scholar] [CrossRef]
  183. González, N.; García, T.; Morant, C.; Barrio, R. Fine-Tuning Intrinsic and Doped Hydrogenated Amorphous Silicon Thin-Film Anodes Deposited by PECVD to Enhance Capacity and Stability in Lithium-Ion Batteries. Nanomaterials 2024, 14, 204. [Google Scholar] [CrossRef] [PubMed]
  184. Puig-Pey González, J.; Lamure, A.; Senocq, F. Polyimide (PI) Films by Chemical Vapor Deposition (CVD): Novel Design, Experiments and Characterization. Surf. Coat. Technol. 2007, 201, 9437–9441. [Google Scholar] [CrossRef]
  185. Kwak, W.; Kim, R.; Lee, J.; Park, H.; Ha, J.; Choi, J. Chemical Vapor Deposition Carbon Coating of SiOx Anode for Li-Ion Batteries: Significance of Carbon Precursor Selection and Deposition Temperature. ACS Omega 2025, 10, 2553–2560. [Google Scholar] [CrossRef]
  186. Keller, C.; Djezzar, Y.; Wang, J.; Karuppiah, S.; Lapertot, G.; Haon, C.; Chenevier, P. Easy Diameter Tuning of Silicon Nanowires with Low-Cost SnO2-Catalyzed Growth for Lithium-Ion Batteries. Nanomaterials 2022, 12, 2601. [Google Scholar] [CrossRef]
  187. Wang, L.; Li, N.; Chen, H.-S.; Song, W.-L. Influence of Carbon Sources on the Performance of Carbon-Coated Nano-Silicon. Chin. Phys. B 2023, 32, 108201. [Google Scholar] [CrossRef]
  188. George, S.M. Atomic Layer Deposition: An Overview. Chem. Rev. 2010, 110, 111–131. [Google Scholar] [CrossRef]
  189. Das, S.; Brennhagen, A.; Cavallo, C.; Killi, V.A.-L.K.; Jensen, I.J.T.; Thøgersen, A.; Mæhlen, J.P.; Lai, S.Y.; Nilsen, O.; Koposov, A.Y. Directing SEI Formation on Si-Based Electrodes Using Atomic Layer Deposition. Chem. Commun. 2024, 60, 15011–15014. [Google Scholar] [CrossRef]
  190. Cao, Y.-Q.; Wang, S.-S.; Liu, C.; Wu, D.; Li, A.-D. Atomic Layer Deposition of ZnO/TiO2 Nanolaminates as Ultra-Long Life Anode Material for Lithium-Ion Batteries. Sci. Rep. 2019, 9, 11526. [Google Scholar] [CrossRef]
  191. Hy, S.; Chen, Y.-H.; Cheng, H.-M.; Pan, C.-J.; Cheng, J.-H.; Rick, J.; Hwang, B.-J. Stabilizing Nanosized Si Anodes with the Synergetic Usage of Atomic Layer Deposition and Electrolyte Additives for Li-Ion Batteries. ACS Appl. Mater. Interfaces 2015, 7, 13801–13807. [Google Scholar] [CrossRef]
  192. Garg, R.; Gonuguntla, S.; Sk, S.; Iqbal, M.S.; Dada, A.O.; Pal, U.; Ahmadipour, M. Sputtering Thin Films: Materials, Applications, Challenges and Future Directions. Curr. Opin. Solid State Mater. Sci. 2024, 330, 103203. [Google Scholar] [CrossRef]
  193. Evshchik, E.; Novikov, D.; Levchenko, A.; Nefedkin, S.; Shikhovtseva, A.V.; Bushkova, O.V.; Dobrovolsky, Y.A. Magnetron Sputtering Silicon Thin Film Electrodes for Lithium-Ion Batteries. Int. J. Electrochem. Sci. 2018, 13, 2860–2874. [Google Scholar] [CrossRef] [PubMed]
  194. Zhang, J.; Hou, Z.-L.; Zhang, X.; Li, C. Si@Cu Composite Anode Material Prepared by Magnetron Sputtering for High-Capacity Lithium-Ion Batteries. Int. J. Hydrogen Energy 2022, 47, 4766–4771. [Google Scholar] [CrossRef]
  195. Cabello, M.; Gucciardi, E.; Herrán, A.; Carriazo, D.; Villaverde, A.; Rojo, T. Towards a High-Power Si@Graphite Anode for Lithium Ion Batteries through a Wet Ball Milling Process. Molecules 2020, 25, 2494. [Google Scholar] [CrossRef]
  196. Koraag, P.Y.E.; Firdaus, A.M.; Hawari, N.H.; Refino, A.D.; Dempwolf, W.; Iskandar, F.; Peiner, E.; Wasisto, H.S.; Sumboja, A. Covalently Bonded Ball-Milled Silicon/CNT Nanocomposite as Lithium-Ion Battery Anode Material. Batteries 2022, 8, 165. [Google Scholar] [CrossRef]
  197. Andersen, H.F.; Foss, C.E.L.; Voje, J.; Tronstad, R.; Mokkelbost, T.; Vullum, P.E.; Ulvestad, A.; Kirkengen, M.; Mæhlen, J.P. Silicon–Carbon Composite Anodes from Industrial Battery Grade Silicon. Sci. Rep. 2019, 9, 14814. [Google Scholar] [CrossRef] [PubMed]
  198. Qian, L.; Lan, J.-L.; Xue, M.; Yu, Y.; Yang, X. Two-Step Ball-Milling Synthesis of a Si/SiOx/C Composite Electrode for Lithium Ion Batteries with Excellent Long-Term Cycling Stability. RSC Adv. 2017, 7, 36697–36704. [Google Scholar] [CrossRef]
  199. Sarwat, S.G. Contamination in Wet-Ball Milling. Powder Metall. 2017, 60, 267–272. [Google Scholar] [CrossRef]
  200. Gao, P.; Fan, X.; Sun, D.; Zeng, G.; Wang, Q.; Wang, Q. Recent Advances in Ball-Milled Materials and Their Applications for Adsorptive Removal of Aqueous Pollutants. Water 2024, 16, 1639. [Google Scholar] [CrossRef]
  201. Zhang, Y.F.; Lu, L.; Yap, S.M. Prediction of the Amount of PCA for Mechanical Milling. J. Mater. Process. Technol. 1999, 89–90, 260–265. [Google Scholar] [CrossRef]
  202. Shang, X.; Wang, X.; Chen, S. Effects of Ball Milling Processing Conditions and Alloy Components on the Synthesis of Cu–Nb and Cu–Mo Alloys. Materials 2019, 12, 1224. [Google Scholar] [CrossRef]
  203. Sun, J.; Li, J.; Ban, B.; Shi, J.; Wang, Q.; Chen, J. A Simple Method to Fabricate Size and Porosity Tunable Si by Al–Si Alloy as Lithium Ion Battery Anode Material. Electrochim. Acta 2020, 345, 136242. [Google Scholar] [CrossRef]
  204. Gomez-Romero, P.; Pokhriyal, A.; Rueda-García, D.; Bengoa, L.N.; González-Gil, R.M. Hybrid Materials: A Metareview. Chem. Mater. 2024, 36, 8–27. [Google Scholar] [CrossRef] [PubMed]
  205. Han, Z.; Wild, J.F.; Chen, J.J.; Yang, Y. Modeling Silane Deposition in Nanoporous Carbon for High-Capacity Si/C Composite Anodes. Energy Mater. Adv. 2024, 5, 0111. [Google Scholar] [CrossRef]
  206. Nguyen, H.T.; Zamfir, M.R.; Duong, L.D.; Lee, Y.H.; Bondavalli, P.; Pribat, D. Alumina-Coated Silicon-Based Nanowire Arrays for High Quality Li-Ion Battery Anodes. J. Mater. Chem. 2012, 22, 24618–24626. [Google Scholar] [CrossRef]
  207. Wang, Y.; Wu, H.; Xiao, F.M.; Tang, R.H.; Sun, T. Synthesis and Electrochemical Performance of Silicon/Carbon Composite Anodes through Ball Milling and Spray Drying Pyrolysis Process. Mater. Sci. Forum 2018, 914, 56–63. [Google Scholar] [CrossRef]
  208. Si, Q.; Hanai, K.; Ichikawa, T.; Phillipps, M.B.; Hirano, A.; Imanishi, N.; Yamamoto, O.; Takeda, Y. Improvement of Cyclic Behavior of a Ball-Milled SiO and Carbon Nanofiber Composite Anode for Lithium-Ion Batteries. J. Power Sources 2011, 196, 9774–9779. [Google Scholar] [CrossRef]
  209. Li, B.; Yao, F.; Bae, J.J.; Chang, J.; Zamfir, M.R.; Le, D.T.; Pham, D.T.; Yue, H.; Lee, Y.H. Hollow Carbon Nanospheres/Silicon/Alumina Core–Shell Film as an Anode for Lithium-Ion Batteries. Sci. Rep. 2015, 5, 7659. [Google Scholar] [CrossRef]
  210. Morand, L.; Iraki, T.; Dornheim, J.; Sandfeld, S.; Link, N.; Helm, D. Machine Learning for Structure-Guided Materials and Process Design. Mater. Des. 2024, 248, 113453. [Google Scholar] [CrossRef]
  211. Qin, G.; Jia, Z.; Sun, S.; Wu, H.; Hu, K.; Liu, D.; Gao, Y.; Chen, J. Carbon-Coated Si Nanosheets as Anode Materials for High-Performance Lithium-Ion Batteries. ACS Appl. Nano Mater. 2024, 7, 7595–7604. [Google Scholar] [CrossRef]
  212. Saint, J.; Morcrette, M.; Larcher, D.; Laffont, L.; Beattie, S.; Pérès, J.-P.; Talaga, D.; Couzi, M.; Tarascon, J.-M. Towards a Fundamental Understanding of the Improved Electrochemical Performance of Silicon–Carbon Composites. Adv. Funct. Mater. 2007, 17, 1765–1774. [Google Scholar] [CrossRef]
  213. John, J.; Gangaja, B.; Nair, S.V.; Santhanagopalan, D. Conformal Coating of TiO2 Shell on Silicon Nanoparticles for Improved Electrochemical Performance in Li-Ion Battery Applications. Electrochim. Acta 2017, 235, 191–199. [Google Scholar] [CrossRef]
  214. Basak, S.; Tavabi, A.H.; Dzieciol, K.; Migunov, V.; Arszelewska, V.; Tempel, H.; Kungl, H.; Kelder, E.M.; Wagemaker, M.; George, C.; et al. Operando Transmission Electron Microscopy of Battery Cycling: Thickness Dependent Breaking of TiO2 Coating on Si/SiO2 Nanoparticles. Chem. Commun. 2022, 58, 3130–3133. [Google Scholar] [CrossRef] [PubMed]
  215. Chabot, V.; Feng, K.; Park, H.W.; Hassan, F.M.; Elsayed, A.R.; Yu, A.; Xiao, X.; Chen, Z. Graphene Wrapped Silicon Nanocomposites for Enhanced Electrochemical Performance in Lithium Ion Batteries. Electrochim. Acta 2014, 130, 127–134. [Google Scholar] [CrossRef]
  216. Costa, A.L.A.; Liebertseder, M.; Gambaryan-Roisman, T.; Esken, D.; Menzel, F. A Comparative Study of Dry-Coated Fumed Metal Oxides for Enhanced Cycling Performance of SiOx/C Anodes in Lithium-Ion Batteries. Electrochem. Commun. 2025, 177, 107941. [Google Scholar] [CrossRef]
  217. Fan, X.-M.; Zhang, X.-H.; Hu, G.-R.; Zhang, B.; He, Z.-J.; Li, Y.-J.; Zheng, J.-C. Single-Walled Carbon Nanotube as Conductive Additive for SiO/C Composite Electrodes in Pouch-Type Lithium-Ion Batteries. Ionics 2019, 25, 5805–5813. [Google Scholar] [CrossRef]
  218. Zhang, X.; Weng, J.; Ye, C.; Liu, M.; Wang, C.; Wu, S.; Tong, Q.; Zhu, M.; Gao, F. Strategies for Controlling or Releasing the Influence Due to the Volume Expansion of Silicon inside Si−C Composite Anode for High-Performance Lithium-Ion Batteries. Materials 2022, 15, 4264. [Google Scholar] [CrossRef]
  219. Shao, Y.; Jin, Z.; Li, J.; Meng, Y.; Huang, X. Evaluation of the Electrochemical and Expansion Performances of the Sn-Si/Graphite Composite Electrode for the Industrial Use. Energy Mater. 2022, 2, 200004. [Google Scholar] [CrossRef]
  220. Yu, Z.; Cui, L.; Zhong, B.; Qu, G. Research Progress on the Structural Design and Optimization of Silicon Anodes for Lithium-Ion Batteries: A Mini-Review. Coatings 2023, 13, 1502. [Google Scholar] [CrossRef]
  221. Duan, H.; Xu, H.; Wu, Q.; Zhu, L.; Zhang, Y.; Yin, B.; He, H. Silicon/Graphite/Amorphous Carbon as Anode Materials for Lithium Secondary Batteries. Molecules 2023, 28, 464. [Google Scholar] [CrossRef]
  222. Yamamoto, M.; Takatsu, M.; Okuno, R.; Kato, A.; Takahashi, M. Nanoporous Silicon Fiber Networks in a Composite Anode for All-Solid-State Batteries with Superior Cycling Performance. Sci. Rep. 2023, 13, 17051. [Google Scholar] [CrossRef]
  223. Jia, H.; Li, X.; Song, J.; Zhang, X.; Luo, L.; He, Y.; Li, B.; Cai, Y.; Hu, S.; Xiao, X.; et al. Hierarchical Porous Silicon Structures with Extraordinary Mechanical Strength as High-Performance Lithium-Ion Battery Anodes. Nat. Commun. 2020, 11, 1474. [Google Scholar] [CrossRef] [PubMed]
  224. Ma, Y.; Zhang, Q.; Shi, L.; Wang, Y.; Wang, T.; Liu, D.; He, D. Low-Cost Micron-Scale Silicon-Based Anodes for High-Performance Lithium-Ion Batteries. ACS Appl. Energy Mater. 2023, 6, 7545–7555. [Google Scholar] [CrossRef]
  225. Maddipatla, R.; Loka, C.; Choi, W.J.; Lee, K.-S. Nanocomposite of Si/C Anode Material Prepared by Hybrid Process of High-Energy Mechanical Milling and Carbonization for Li-Ion Secondary Batteries. Appl. Sci. 2018, 8, 2140. [Google Scholar] [CrossRef]
  226. Fang, Z.; Zhou, P.; Tian, Y.; Fang, H.; Zhang, Q. One-Step Chemical Vapor Deposition Synthesis of Si NWs@C Core/Shell Anodes without Additional Catalysts by the Oxide-Assisted Growth Mechanism for Lithium-Ion Batteries. Dalton Trans. 2024, 53, 9052–9061. [Google Scholar] [CrossRef]
  227. Du, C.; Chen, M.; Wang, L.; Yin, G. Covalently-Functionalizing Synthesis of Si@C Core–Shell Nanocomposites as High-Capacity Anode Materials for Lithium-Ion Batteries. J. Mater. Chem. 2011, 21, 15692–15697. [Google Scholar] [CrossRef]
  228. Shen, B.H.; Wang, S.; Tenhaeff, W.E. Ultrathin Conformal Polycyclosiloxane Films to Improve Silicon Cycling Stability. Sci. Adv. 2019, 5, eaaw4856. [Google Scholar] [CrossRef]
  229. Peng, B.; Bao, W.; Sun, K.; Xiao, J. Enhancing Silicon Anode Performance in Lithium-Ion Batteries Through Hybrid Artificial SEI Layer and Prelithiation. Nanomaterials 2025, 15, 690. [Google Scholar] [CrossRef] [PubMed]
  230. Zhang, X.; Wang, D.; Qiu, X.; Ma, Y.; Kong, D.; Müllen, K.; Li, X.; Zhi, L. Stable High-Capacity and High-Rate Silicon-Based Lithium Battery Anodes upon Two-Dimensional Covalent Encapsulation. Nat. Commun. 2020, 11, 3826. [Google Scholar] [CrossRef] [PubMed]
  231. Liu, C.; Lombardo, T.; Xu, J.; Ngandjong, A.C.; Franco, A.A. An Experimentally-Validated 3D Electrochemical Model Revealing Electrode Manufacturing Parameters’ Effects on Battery Performance. Energy Storage Mater. 2023, 54, 156–163. [Google Scholar] [CrossRef]
Figure 2. Schematic illustration of the synthesis for Si/Cu3Si-based composite powder. Reproduced from ref. [24], licensed under CC BY 4.0 (https://creativecommons.org/licenses/by/4.0/). Accessed on 30 May 2025.
Figure 2. Schematic illustration of the synthesis for Si/Cu3Si-based composite powder. Reproduced from ref. [24], licensed under CC BY 4.0 (https://creativecommons.org/licenses/by/4.0/). Accessed on 30 May 2025.
Batteries 11 00346 g002
Figure 3. Schematic illustration of lithiation and delithiation processes of YS Si@10mC and YS Si@50mC electrodes. Reproduced from ref. [29] with permission from Elsevier. Copyright © 2015 Elsevier.
Figure 3. Schematic illustration of lithiation and delithiation processes of YS Si@10mC and YS Si@50mC electrodes. Reproduced from ref. [29] with permission from Elsevier. Copyright © 2015 Elsevier.
Batteries 11 00346 g003
Figure 4. Schematic diagram of different structures of Si/C composite anodes for lithium-ion batteries. Reproduced from ref. [41], licensed under CC BY 4.0 (https://creativecommons.org/licenses/by/4.0/). Accessed on 30 May 2025.
Figure 4. Schematic diagram of different structures of Si/C composite anodes for lithium-ion batteries. Reproduced from ref. [41], licensed under CC BY 4.0 (https://creativecommons.org/licenses/by/4.0/). Accessed on 30 May 2025.
Batteries 11 00346 g004
Figure 5. (a) Schematic illustration of the synthesis process of cpDOPA-crGO–Si and diagram of the BUCHI-290 mini spray-dryer used. (① N2 flow in ② spray nozzle ③ drying chamber ④ cyclone ⑤ product vessel ⑥ filter.) (b) Long-term cycling performance. Figure 1 and Figure 3. Reproduced from ref. [63] licensed under CC BY-NC-ND 4.0 (https://creativecommons.org/licenses/by-nc-nd/4.0/). No modifications were made. Copyright © 2021 American Chemical Society. Accessed on 30 May 2025.
Figure 5. (a) Schematic illustration of the synthesis process of cpDOPA-crGO–Si and diagram of the BUCHI-290 mini spray-dryer used. (① N2 flow in ② spray nozzle ③ drying chamber ④ cyclone ⑤ product vessel ⑥ filter.) (b) Long-term cycling performance. Figure 1 and Figure 3. Reproduced from ref. [63] licensed under CC BY-NC-ND 4.0 (https://creativecommons.org/licenses/by-nc-nd/4.0/). No modifications were made. Copyright © 2021 American Chemical Society. Accessed on 30 May 2025.
Batteries 11 00346 g005
Figure 6. (a) Schematic illustration for the synthesis of core–shell structured Si@Fe2O3/C composite. (b) Cycling performance. Reproduced from ref. [81] with permission from Elsevier. Copyright © 2019 Elsevier B.V.
Figure 6. (a) Schematic illustration for the synthesis of core–shell structured Si@Fe2O3/C composite. (b) Cycling performance. Reproduced from ref. [81] with permission from Elsevier. Copyright © 2019 Elsevier B.V.
Batteries 11 00346 g006
Figure 7. (a) Schematic diagram of the process for synthesizing BA and AA electrodes. (b) Comparison of capacities among BA, AA electrode, and blank sample [98].
Figure 7. (a) Schematic diagram of the process for synthesizing BA and AA electrodes. (b) Comparison of capacities among BA, AA electrode, and blank sample [98].
Batteries 11 00346 g007
Figure 8. (a) Schematic illustration of the preparation of PPy@PHSi nanocomposite. (b) Their cycling and rate performance. Reproduced with permission from ref. [119]. Copyright 2014 Wiley-VCH.
Figure 8. (a) Schematic illustration of the preparation of PPy@PHSi nanocomposite. (b) Their cycling and rate performance. Reproduced with permission from ref. [119]. Copyright 2014 Wiley-VCH.
Batteries 11 00346 g008
Figure 9. (a) Schematic illustration of the preparation process of Si@void@C anode material by the presented templateless method. (b) TEM images of bare Si (left) and Si@void@C (right). (c) Cycling performance of H-Si@void@C, M-Si@void@C, and L-Si@void@C anodes at 2.0 A/g after activation at 0.1 A/g for two cycles, along with the corresponding Coulombic efficiency of the M-Si@void@C anode. Reprinted (adapted) with permission from ref. [133]. Copyright 2019 American Chemical Society.
Figure 9. (a) Schematic illustration of the preparation process of Si@void@C anode material by the presented templateless method. (b) TEM images of bare Si (left) and Si@void@C (right). (c) Cycling performance of H-Si@void@C, M-Si@void@C, and L-Si@void@C anodes at 2.0 A/g after activation at 0.1 A/g for two cycles, along with the corresponding Coulombic efficiency of the M-Si@void@C anode. Reprinted (adapted) with permission from ref. [133]. Copyright 2019 American Chemical Society.
Batteries 11 00346 g009
Figure 10. (a) Schematic fabrication process for multilayered Si/RGO nanostructures. (b) Schematic representation of structural change (not to scale) of a multilayered Si/RGO anode before and after lithiation. Top and bottom RGO films are represented by yellow and light blue. Si NPs and Li-Si NPs are shown as magenta and cyan spheres. (c) Electrochemical characteristics of layered Si/RGO nanostructures. Left axis: capacity retention for Si/RGO and Si anode electrodes at 0.05 C-rate for the first two cycles and then 1 C for subsequent cycles; right axis: Coulombic efficiency of the Si/RGO electrode. Reproduced with permission from ref. [144]. Copyright 2013 Wiley-VCH.
Figure 10. (a) Schematic fabrication process for multilayered Si/RGO nanostructures. (b) Schematic representation of structural change (not to scale) of a multilayered Si/RGO anode before and after lithiation. Top and bottom RGO films are represented by yellow and light blue. Si NPs and Li-Si NPs are shown as magenta and cyan spheres. (c) Electrochemical characteristics of layered Si/RGO nanostructures. Left axis: capacity retention for Si/RGO and Si anode electrodes at 0.05 C-rate for the first two cycles and then 1 C for subsequent cycles; right axis: Coulombic efficiency of the Si/RGO electrode. Reproduced with permission from ref. [144]. Copyright 2013 Wiley-VCH.
Batteries 11 00346 g010
Figure 11. (a) Schematic illustration of preparation and structural diagram of the CHSP. (b) Rate performance of CHSP and nano-Si. Reproduced with permission from ref. [148]. Copyright 2023 Elsevier.
Figure 11. (a) Schematic illustration of preparation and structural diagram of the CHSP. (b) Rate performance of CHSP and nano-Si. Reproduced with permission from ref. [148]. Copyright 2023 Elsevier.
Batteries 11 00346 g011
Figure 12. (a) A schematic illustration of the synthetic process of the Si@HC yolk–shell structure. (b) TEM and STEM images of Si@HC-1 before (①, ②) and after HF etching (③, ④), and the elemental mapping of Si@HC-1. (c) Cycling performances at 100 mA/g (left), and rate performances at different current densities of Si@HC-0.5, Si@HC-1 and Si@HC-2 (right). Reproduced with permission from ref. [170]. Copyright 2020 Elsevier.
Figure 12. (a) A schematic illustration of the synthetic process of the Si@HC yolk–shell structure. (b) TEM and STEM images of Si@HC-1 before (①, ②) and after HF etching (③, ④), and the elemental mapping of Si@HC-1. (c) Cycling performances at 100 mA/g (left), and rate performances at different current densities of Si@HC-0.5, Si@HC-1 and Si@HC-2 (right). Reproduced with permission from ref. [170]. Copyright 2020 Elsevier.
Batteries 11 00346 g012
Figure 13. (a) Schematic of Si and Si@Cu composite material prepared by magnetron sputtering. Electrochemical performance of Si electrode, Si@Cu composite electrode and Si@Cu@C composite electrode (b) cyclic performance and (c) rate performance. Reproduced with permission from ref. [194]. Copyright 2022 Elsevier.
Figure 13. (a) Schematic of Si and Si@Cu composite material prepared by magnetron sputtering. Electrochemical performance of Si electrode, Si@Cu composite electrode and Si@Cu@C composite electrode (b) cyclic performance and (c) rate performance. Reproduced with permission from ref. [194]. Copyright 2022 Elsevier.
Batteries 11 00346 g013
Figure 14. (a) Cycling performance of the cells with Si30h and Si30h/CNT electrodes, (b) rate performance, and (c) Nyquist plot with the fitted equivalent circuit of the cells with Si30h and Si30h/CNT electrodes. (d) SEM (①) and TEM (②) images of Si30h; SEM (③) and TEM (④) images of Si30h/CNT. Reproduced from ref. [196]. © 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/). Accessed on 30 May 2025.
Figure 14. (a) Cycling performance of the cells with Si30h and Si30h/CNT electrodes, (b) rate performance, and (c) Nyquist plot with the fitted equivalent circuit of the cells with Si30h and Si30h/CNT electrodes. (d) SEM (①) and TEM (②) images of Si30h; SEM (③) and TEM (④) images of Si30h/CNT. Reproduced from ref. [196]. © 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/). Accessed on 30 May 2025.
Batteries 11 00346 g014
Figure 15. Electrochemical performance of different silicon anodes (Si/SiO2 with no coating, with 5 nm TiO2 coating and with 10 nm TiO2 coating) at different C rates. The first 5 cycles are at C/10, at C/5 and at C/2 rates. Reproduced from ref. [214]. Licensed under CC BY 3.0.
Figure 15. Electrochemical performance of different silicon anodes (Si/SiO2 with no coating, with 5 nm TiO2 coating and with 10 nm TiO2 coating) at different C rates. The first 5 cycles are at C/10, at C/5 and at C/2 rates. Reproduced from ref. [214]. Licensed under CC BY 3.0.
Batteries 11 00346 g015
Figure 16. Comparison of discharge capacity and Coulombic efficiency between graphite and composite anodes (Sn-Si hybrid/C = 1:9) in half-cell. Reproduced from ref. [222]. Licensed under CC BY 4.0.
Figure 16. Comparison of discharge capacity and Coulombic efficiency between graphite and composite anodes (Sn-Si hybrid/C = 1:9) in half-cell. Reproduced from ref. [222]. Licensed under CC BY 4.0.
Batteries 11 00346 g016
Figure 17. The initial charge and discharge profiles of Si/G/C and Si/C anodes at 0.1 A/g. Reproduced from ref. [221]. Licensed under CC BY 4.0.
Figure 17. The initial charge and discharge profiles of Si/G/C and Si/C anodes at 0.1 A/g. Reproduced from ref. [221]. Licensed under CC BY 4.0.
Batteries 11 00346 g017
Figure 18. Preparation process and mechanism of the HPC/Si@C composite. Reproduced with permission from ref. [64]. Copyright © 2024 Elsevier.
Figure 18. Preparation process and mechanism of the HPC/Si@C composite. Reproduced with permission from ref. [64]. Copyright © 2024 Elsevier.
Batteries 11 00346 g018
Figure 19. Schematic illustration of fabricating porous Si (P–Si) and carbon-coated porous Si (P–Si@C) microparticles. Reprinted with permission from ref. [224]. Copyright 2023 American Chemical Society.
Figure 19. Schematic illustration of fabricating porous Si (P–Si) and carbon-coated porous Si (P–Si@C) microparticles. Reprinted with permission from ref. [224]. Copyright 2023 American Chemical Society.
Batteries 11 00346 g019
Table 1. Electrochemical performance summary of silicon-based hybrid anode systems reported in the references of Section 3.1.1. Only numerical data explicitly reported in the cited studies are included.
Table 1. Electrochemical performance summary of silicon-based hybrid anode systems reported in the references of Section 3.1.1. Only numerical data explicitly reported in the cited studies are included.
Si AnodesCurrent Density (A/g)Cycle NumberRemaining Capacity (mAh/g)Source
(1st Author, Year, [Ref. No.])
Silicon-Carbon
Si/C0.1100941.1Zhang, 2023, [42]
Si/C0.5100605.43Li, 2023, [43]
Si/C2.04001283Ma, 2023, [44]
Si@C@void@C0.5501366Xie, 2017, [45]
Meso-Si/C1.01000990Xu, 2016, [46]
Si/graphite/C0.5300~400Wang, 2016, [47]
M-pSi@C1.02501702Su, 2023, [48]
Porous Si/C nanotubes0.22001300Chen, 2017, [49]
Si/C/graphene0.2100760Wu, 2015, [50]
Si/rGO0.11001433Liu, 2015, [51]
Si/CNTs5.01000612.3Guo, 2024, [53]
Si/CNTs10100800Gohier, 2012, [59]
Si@(POH-AOCNTs)1.05001195.8Liu, 2024, [60]
Si@CNFs1.0100658.1Zhang, 2025, [61]
cpDOPA-crGO-Si1.02001038She, 2021, [63]
HPC/Si@C2.0500358Cao, 2024, [64]
SiO2@NPC Y.S.0.1300705Min, 2022, [66]
Table 2. Electrochemical performance summary of silicon-based hybrid anode systems reported in the references of Section 3.1.2. Only numerical data explicitly reported in the cited studies are included.
Table 2. Electrochemical performance summary of silicon-based hybrid anode systems reported in the references of Section 3.1.2. Only numerical data explicitly reported in the cited studies are included.
Si Anodes.Current Density (A/g)Cycle NumberRemaining Capacity (mAh/g)Source
(1st Author, Year, [Ref. No.])
Silicon-Metal Oxide
Porous Si/Fe2O3 0.2100697.2Chen, 2020, [70]
Si@SnO20.22001926Zhu, 2022, [72]
Si@a-TiO20.422001720Yang, 2017, [74]
Si/TiO20.11001010.7Li, 2016, [75]
TiO2/SiNWs0.1100~1700Lotfabad, 2013, [77]
SiO/Fe2O30.16501335Zhou, 2013, [79]
Si/Fe2O3@rGO0.22001744.5Yan, 2021, [80]
Si@Fe2O3/C1.0300680.7Wang, 2019, [81]
Si-SnO2-graphene0.120350Li, 2017, [85]
TiGL@Si1.03501200Huertas, 2022, [89]
Table 3. Electrochemical performance summary of silicon-based hybrid anode systems reported in the references of Section 3.1.3. Only numerical data explicitly reported in the cited studies are included.
Table 3. Electrochemical performance summary of silicon-based hybrid anode systems reported in the references of Section 3.1.3. Only numerical data explicitly reported in the cited studies are included.
Si AnodesCurrent Density (A/g)Cycle NumberRemaining Capacity (mAh/g)Source
(1st Author, Year, [Ref. No.])
Silicon-2D Material
Si@Ti3C2 Mxene0.2150188Kong, 2018, [91]
Si@Ti3C2 Mxene1.02001342.8Yang, 2020, [92]
nSi/MX-N1.5701106Zhang, 2019, [93]
Si/Mxene0.21002118Tian, 2019, [94]
MoS2@Si0.13001223Marriam, 2023, [95]
Si/d-Ti3C20.52001130Zhu, 2019, [97]
nmSi-NH2/MXene1.0100929.5Kong, 2024, [98]
Si/MoS2-G0.290923Kawade, 2019, [103]
Table 4. Electrochemical performance summary of silicon-based hybrid anode systems reported in the references of Section 3.1.4. Only numerical data explicitly reported in the cited studies are included.
Table 4. Electrochemical performance summary of silicon-based hybrid anode systems reported in the references of Section 3.1.4. Only numerical data explicitly reported in the cited studies are included.
Si AnodesCurrent Density (A/g)Cycle NumberRemaining Capacity (mAh/g)Source
(1st Author, Year, [Ref. No.])
Silicon-Conductive Polymer
SiNP/PANi1.010001200Wu, 2013, [108]
Si-oxalic acid-PANI1.01000610Zhou, 2020, [109]
Si@Ppy1.05001047Zhang, 2024, [110]
Si/PANI1.0100545.3Tu, 2013, [113]
Si/PANi/C0.12001470Mu, 2019, [114]
Si/PANi0.1251840Cai, 2010, [116]
Si@PEDOT0.51001439.8Li, 2021, [121]
PEDOT:PSS/SiNP1.01001950Higgins, 2016, [122]
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

Jang, J.; Kwon, T. Various Technologies to Mitigate Volume Expansion of Silicon Anode Materials in Lithium-Ion Batteries. Batteries 2025, 11, 346. https://doi.org/10.3390/batteries11090346

AMA Style

Jang J, Kwon T. Various Technologies to Mitigate Volume Expansion of Silicon Anode Materials in Lithium-Ion Batteries. Batteries. 2025; 11(9):346. https://doi.org/10.3390/batteries11090346

Chicago/Turabian Style

Jang, Jihun, and Taegyun Kwon. 2025. "Various Technologies to Mitigate Volume Expansion of Silicon Anode Materials in Lithium-Ion Batteries" Batteries 11, no. 9: 346. https://doi.org/10.3390/batteries11090346

APA Style

Jang, J., & Kwon, T. (2025). Various Technologies to Mitigate Volume Expansion of Silicon Anode Materials in Lithium-Ion Batteries. Batteries, 11(9), 346. https://doi.org/10.3390/batteries11090346

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