Next Article in Journal
Scaling, Leakage Current Suppression, and Simulation of Carbon Nanotube Field-Effect Transistors
Previous Article in Journal
Electrical Properties of Carbon Nanotubes: From Individual to Assemblies
Previous Article in Special Issue
The Air Stability of Sodium Layered Oxide NaTMO2 (100) Surface Investigated via DFT Calculations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in the Application of VO2 for Electrochemical Energy Storage

College of Chemistry, Chemical Engineering and Resource Utilization, Northeast Forestry University, Harbin 150040, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(15), 1167; https://doi.org/10.3390/nano15151167
Submission received: 29 June 2025 / Revised: 20 July 2025 / Accepted: 26 July 2025 / Published: 28 July 2025
(This article belongs to the Special Issue Nanostructured Materials for Energy Storage)

Abstract

Energy storage technology is crucial for addressing the intermittency of renewable energy sources and plays a key role in power systems and electronic devices. In the field of energy storage systems, multivalent vanadium-based oxides have attracted widespread attention. Among these, vanadium dioxide (VO2) is distinguished by its key advantages, including high theoretical capacity, low cost, and strong structural designability. The diverse crystalline structures and plentiful natural reserves of VO2 offer a favorable foundation for facilitating charge transfer and regulating storage behavior during energy storage processes. This mini review provides an overview of the latest progress in VO2-based materials for energy storage applications, specifically highlighting their roles in lithium-ion batteries, zinc-ion batteries, photoassisted batteries, and supercapacitors. Particular attention is given to their electrochemical properties, structural integrity, and prospects for development. Additionally, it explores future development directions to offer theoretical insights and strategic guidance for ongoing research and industrial application of VO2.

Graphical Abstract

1. Introduction

Energy is a fundamental driving force of societal development and serves as the material foundation of daily human life. Energy storage devices not only absorb surplus energy to prevent waste but also release it during critical periods, thereby ensuring the safe and efficient use of energy. In modern society, electricity has penetrated all aspects of daily life, ranging from lighting and household appliances to transportation. Batteries and capacitors play indispensable roles in electronic devices and power systems. Batteries provide high energy density for sustained operation [1], whereas supercapacitors deliver rapid charge/discharge capabilities to enable instantaneous power delivery [2]. Collectively, they form fundamental components of modern electronics and energy infrastructure.
To design high-performance energy storage devices, a variety of functional materials have been developed, including transition metal oxides (e.g., VO2, V2O5 [3,4], MnO2 [5,6,7]) and carbon-based materials (e.g., graphene [8], carbon nanotubes [9,10]), MXenes [11], and transition metal chalcogenides (e.g., MoS2 [12]). These materials exhibit excellent electrochemical properties, tunable structures, and strong photoelectric responses, and have been widely employed in high-capacity, high-efficiency, and long-lifespan lithium-ion batteries, zinc-ion batteries, and other energy storage systems.
Among these materials, VO2, a representative vanadium-based oxide, has attracted considerable attention because of its high theoretical capacity and excellent chemical stability. Owing to its diverse crystalline phases and abundant availability, VO2 can efficiently regulate charge transport and storage during electrochemical cycling, demonstrating great potential in the development of multifunctional and high-performance energy storage systems. This paper systematically reviews recent advances in VO2-based materials for energy storage technologies, with a particular focus on their applications in lithium-ion batteries, zinc-ion batteries, photoassisted batteries, and supercapacitors. The review is based primarily on high-quality publications from the past three years, aiming to capture the most up-to-date and impactful progress in the field. It also discusses future research directions, providing a theoretical foundation for ongoing and future developments.

2. Crystalline Structure and Synthesis of VO2

2.1. Crystalline Structure of VO2

VO2 has a variety of crystalline phases, making it a promising candidate for diverse applications. It exists in several polymorphic forms, including the tetragonal rutile phase VO2 (R), the monoclinic phase VO2 (M), the triclinic phase VO2 (T), and several metastable phases, such as VO2 (A), VO2 (B), VO2 (C), VO2 (D), VO2 (P), and VO2 (N) [13,14,15,16,17,18,19,20,21]. This review highlights four common VO2 polymorphs: VO2 (B) [17], VO2 (D) [19], VO2 (M) [14], and VO2 (R) [13].
VO2 (B) is a metastable monoclinic phase with lattice parameters of a = 12.09 Å, b = 3.702 Å, c = 6.433 Å, and β = 106.6°, and it belongs to the space group C 2/m. The VO2 (B) lattice comprises two layers of distorted VO6 octahedra aligned along the b-axis (Figure 1a,b). Each vanadium atom is situated at the center of an octahedron and is surrounded by six oxygen atoms, forming VO6 units. These [VO6] octahedra form double chains and are connected into a two-dimensional layered structure via corner-sharing oxygen atoms [17]. The structure is characterized by interlayer voids and tunnels, which provide effective channels for ion intercalation and deintercalation. Moreover, this architecture can effectively mitigate volume expansion during electrochemical cycling.
VO2 (D) is another metastable monoclinic phase with space group P 2/c and lattice parameters of a = 4.60 Å, b = 5.68 Å, c = 4.91 Å, and β = 89.39° [19]. Its structure comprises [V(1)O6] octahedra that share edges to form zigzag chains (Figure 1c,d). These chains are interconnected with those formed by [V(2)O6] octahedra through corner-sharing oxygen atoms, resulting in a three-dimensional framework.
VO2 (M) is classified as a monoclinic phase with space group P21/c and lattice parameters a = 5.75 Å, b = 4.53 Å, c = 5.38 Å, α = γ = 90°, and β = 122.60°. In this configuration, vanadium atoms form alternating zigzag chains along the x-axis (Figure 1e,f). These chains display alternating V–V bond lengths of 2.65 Å and 3.12 Å, indicating the occurrence of V–V dimerization [14]. The structure consists of twisted VO6 octahedra connected through coplanar edges, and the chains are further bridged by corner-sharing oxygen atoms to form a robust three-dimensional framework.
VO2 (R) crystallizes in a tetragonal rutile structure with a space group of P42/mnm and lattice parameters a = b = 4.55 Å, c = 2.88 Å, and β = 90° [13]. In this structure, V4+ ions occupy the corners and body center of the unit cell and are coordinated by six O2− ions to form nearly regular [VO6] octahedra (Figure 1g,h). These octahedra are connected via edge sharing along the c-axis, forming continuous linear chains. Between adjacent chains, the octahedra are linked through corner sharing, resulting in a highly symmetric and thermally stable metallic framework.
In summary, each VO2 polymorph exhibits distinct structural characteristics that influence its suitability for energy storage [22,23]. VO2 (B), with its open-layered monoclinic structure and interconnected tunnels, offers fast ion diffusion and high specific capacity, making it highly promising for battery and supercapacitor applications. VO2 (D) shares similar diffusion pathways and electrochemical activities but has been less studied due to phase instability and synthesis challenges. VO2 (M), the thermodynamically stable phase at room temperature, provides good structural integrity and moderate performance but suffers from a dense framework that limits rate capability. VO2 (R), despite its high electronic conductivity and symmetrical rutile structure, is thermally unstable under ambient conditions and readily reverts to VO2 (M), thus limiting its practical applicability. These differences underscore the importance of phase selection and structural tailoring in optimizing VO2-based materials for energy storage devices.

2.2. Synthetic Method of VO2

In the field of electrochemical energy storage, VO2 has gained significant attention because of its excellent ion mobility and high specific capacity. The development and optimization of its synthesis methods have remained key areas of focus for both academia and industry. This mini review briefly outlines the main synthesis strategies for VO2 used in energy storage applications, including hydrothermal synthesis, pyrolysis, electrodeposition, pulsed laser deposition (PLD), and other methods.

2.2.1. Hydrothermal Methods

The hydrothermal method is one of the most widely used techniques for the synthesis of VO2 materials. It involves a dissolution–crystallization process in an aqueous solution under high-temperature and high-pressure conditions. For VO2 hydrothermal synthesis, pentavalent vanadium sources such as V2O5 and ammonium metavanadate are commonly employed, whereas oxalic acid and formic acid serve as reducing agents. By adjusting the synthesis parameters, such as the reaction temperature, duration, and reducing agent type, nanomaterials of various sizes and morphologies, including nanosheets, nanorods, nanoflowers, and nanoparticles, can be obtained.
Hou et al. synthesized homogeneous VO2 (B) nanorods from V2O5 and oxalic acid via a facile and rapid hydrothermal method [24]. A heterogeneous structure (A-VO2) with a crystalline core and an amorphous shell layer was further constructed by disorder–order engineering of VO2 (B) nanorods through a simple reduction treatment in NaBH4 solution (Figure 2a). The tunable amorphous layer introduces abundant oxygen vacancies, thereby facilitating ion and electron transport and enhancing structural stability. Pan et al. successfully synthesized VO2 (D) hollow microspheres with yolk-shell, multishell, and monoshell structures via a one-step template-free solvothermal approach [25]. The internal structure can be controlled by adjusting the reaction time and precursor concentration (Figure 2b). The hollow structure remained intact after calcination, and the resulting VO2 (D) hollow composite structure exhibited high stability, mitigating the agglomeration and dissolution issues typical of conventional solid particles and thereby enhancing ion transport. Liu et al. synthesized nsutite-type VO2 microcrystals composed of nanosheets using NH4VO3 as the precursor and thioacetamide as the reductant [21]. These microcrystals were then assembled into zinc batteries as cathode materials. The microcrystalline structure collapsed into nanosheets during discharge and reassembled into nanosheets during charging. Furthermore, the nanosheets were converted into nanoplates approximately 100 nm thick during charging, and this reversible structural change significantly improved the cycling stability and ion transport efficiency of the material.
The hydrothermal method operates under mild conditions with precise control over temperature and pressure, enabling high crystallinity, uniform particle size, and controlled morphology. However, it typically involves long reaction times and is less suitable for large-scale production.

2.2.2. Pyrolysis Methods

Pyrolysis refers to the synthesis of VO2 nanostructures by exploiting the thermal instability of precursors such as vanadates, vanadium-based oxides, and their hydrates, which are decomposed through thermal treatment in an anaerobic environment at medium to high temperatures. Thakur et al. successfully prepared VO2 (M) nanopowders with particle sizes < 30 nm through pyrolytic decomposition of the precursor [NH4]5[(VO)6(CO3)4(OH)9]·10H2O [26]. The significant release of gases such as NH3 and CO2 during pyrolysis promotes the nanosizing of VO2 (M) powders. By adjusting the pyrolysis conditions, such as the precursor size, heating rate, and gas flow rate, the stoichiometric ratio (VO2 ± x, x = 0.04–0.07) and crystalline structure (monocrystalline, nanocrystalline, and amorphous) of the products can be tailored.
Jung et al. developed a method to synthesize high-purity monoclinic phase VO2 (M) powders via a mild pyrolysis process without an inert gas atmosphere. VO2 (M) powder can be synthesized within 1 h at a low temperature [27]. Thermal analysis revealed that the pyrolysis temperature should not exceed 253 °C to prevent oxidation to V2O5. The precursor vanadium ethanolate was synthesized and heat-treated for 1 h at 190 °C with an airflow rate of 10 L min−1, resulting in VO2 (M) powders consisting of 20–50 nm spherical nanoparticles that agglomerated into porous nanorods. These nanorods exhibit reversible metal–insulator phase transition properties.
Pyrolysis allows stoichiometric control through thermal decomposition and is scalable for large-scale production. However, high temperatures may cause particle agglomeration and limit morphology control.

2.2.3. Other Methods

The electrodeposition method allows precise control over the morphology and thickness of the deposited layer by regulating the current density and time. It can be directly deposited on complex substrates without the need for additional binders, making it suitable for the development of energy storage electrodes. Lai et al. used a constant-current electrodeposition method to prepare amorphous VO4 nanowires from VOSO4 as a vanadium source, with H2O2 as the oxidizing agent, on carbon cloth [28]. These nanowires were subsequently transformed into uniform, porous VO2 (B) nanowires by annealing and crystallization (Figure 2d). This VO2 (B) phase has many active sites and fast ion diffusion capabilities, leading to high performance in zinc-ion memory devices. However, this VO2 (B) phase can irreversibly convert to V2O5·H2O, causing capacity degradation. To address this, structural design modifications and other approaches are necessary to inhibit phase transition and vanadium solubilization, thereby increasing cycling stability. Xiang et al. developed a simple method for synthesizing high-quality VO2 (M) thin films on polymer substrates by preparing VO2 (M) films through room-temperature PLD followed by annealing at 390 °C in an oxygen atmosphere [29]. The resulting VO2 (M) films exhibit excellent electrical phase transition properties and outstanding optical modulation. By coupling tungsten doping with film strain, the transition temperature of VO2 (M) can be adjusted to room temperature with a doping concentration of only 1.1% while maintaining a high electrical contrast of two orders of magnitude. PLDs allow for precise structural engineering and direct integration with substrates, although their scalability and cost-effectiveness remain limited. Kim et al. proposed a facile and scalable strategy for the first successful synthesis of two-dimensional VO2 (M) nanosheets with a high aspect ratio and high crystallinity through the heat treatment of monolayer V2CTx MXene nanosheets (Figure 2c). The resulting VO2 (M) nanosheets can be directly sprayed onto flexible or nonplanar substrates to form dense and uniform films that exhibit excellent thermochromic properties [30].
In addition, sputtering [31] and electrostatic spinning [32] methods have also been employed for the fabrication of VO2 nanostructures. These techniques offer distinct advantages in regulating nanostructure morphology (e.g., nanowires, nanofilms, and nanofibers) and crystallinity, providing diverse approaches for the controllable preparation of VO2 functional materials. Overall, each method has unique trade-offs in terms of structural control, energy efficiency, and compatibility with device fabrication.
Despite these advancements, many current synthesis routes still face challenges related to high energy consumption, complex processing steps, and limited scalability. Addressing these limitations will require the development of simplified, low-temperature, and continuous synthesis strategies. Solid-state reactions with optimized precursors and hydrothermal processes using recyclable solvents and shorter reaction times are especially promising for facilitating industrial-scale production.
In parallel, VO2 itself has relatively low toxicity and good chemical stability under typical operating conditions, supporting its potential use in both large-scale and wearable energy storage systems [33,34]. Nonetheless, despite the effective suppression of vanadium ion leaching through surface modifications [35], further comprehensive studies on its long-term biocompatibility and environmental impact remain essential to ensure safe and sustainable use.
Figure 2. (a) Schematic diagram of the synthesis process of VO2 and A-VO2 samples. Reproduced with permission [24]. Copyright 2024, Elsevier. (b) Time-dependent structural evolution of VO2 microspheres. Reproduced with permission [25]. Copyright 2013, Wiley-VCH GmbH. (c) Schematic for the fabrication of V2CTx-nanosheet-converted VO2 nanosheets. Reproduced with permission [28]. Copyright 2023, Elsevier. (d) Illustration of the synthetic procedure of VO2/CC. Reproduced with permission [30]. Copyright 2024, Elsevier.
Figure 2. (a) Schematic diagram of the synthesis process of VO2 and A-VO2 samples. Reproduced with permission [24]. Copyright 2024, Elsevier. (b) Time-dependent structural evolution of VO2 microspheres. Reproduced with permission [25]. Copyright 2013, Wiley-VCH GmbH. (c) Schematic for the fabrication of V2CTx-nanosheet-converted VO2 nanosheets. Reproduced with permission [28]. Copyright 2023, Elsevier. (d) Illustration of the synthetic procedure of VO2/CC. Reproduced with permission [30]. Copyright 2024, Elsevier.
Nanomaterials 15 01167 g002

3. Application of VO2 in Electrochemical Energy Storage

Among the various materials explored for energy storage applications, VO2 has attracted considerable attention because of its moderate cost, structural tunability, and well-rounded electrochemical performance (Table 1). Compared with other vanadium oxides (e.g., V2O5, V6O13) and emerging materials such as MXenes (e.g., Ti3C2I2) and MoS2, VO2 offers a more balanced profile in terms of capacity, stability, scalability, and cost [36,37,38,39,40,41]. While V2O5 and V6O13 provide low cost or high capacity, they suffer from poor cycling stability. Ti3C2I2 has excellent conductivity but lacks sufficient capacity and scalability, and MoS2 offers good stability with relatively low energy density. These trade-offs highlight VO2 as a competitive and practical candidate for next-generation rechargeable batteries.
In the following sections, recent advances in VO2-based materials over the past three years are systematically reviewed, with a focus on their applications in lithium-ion batteries, zinc-ion batteries, photoassisted batteries, and supercapacitors.

3.1. Lithium-Ion Batteries

Among various emerging energy storage technologies, lithium-ion batteries are widely regarded as the dominant option because of their high energy density, long cycle life, and well-established industrial infrastructure [42,43]. In recent years, VO2 has attracted considerable attention as a cathode material for high-capacity lithium-ion batteries owing to its unique layered structure and reversible lithium storage capability.
Castro-Pardo et al. systematically investigated the influence of the metal–insulator transition (MIT) of VO2 on its electrochemical performance as a lithium-ion battery cathode [44]. Their findings demonstrated that transitioning VO2 from the monoclinic (M) phase to the rutile (R) phase near 68 °C led to an ~70% increase in specific capacity, significantly enhanced rate performance, and reduced charge-transfer resistance. These improvements originate from a series of structural and electronic changes during the MIT. In the low-temperature monoclinic phase, lithium ions migrate through zigzag diffusion paths formed by distorted V–V chains, which impose high migration energy barriers. Upon transition to the rutile phase, the V–V chains straighten into a linear arrangement, forming direct, low-resistance channels for Li+ diffusion (Figure 3a,b). This structural reconfiguration significantly lowers the diffusion barrier, as evidenced by a tenfold increase in the Li+ diffusion coefficient. Simultaneously, the MIT drastically enhances the electronic conductivity by 3–5 orders of magnitude, ensuring better charge transport throughout the electrode. DFT calculations further show that the rutile phase accommodates a higher lithium filling fraction (0.67 vs. 0.42 per VO2 unit) and offers more energetically favorable insertion sites, facilitating a higher reversible capacity. Together, these synergistic effects of structural alignment and electronic enhancement make the VO2 (R) phase more suitable for fast and stable lithium storage. In summary, the MIT in VO2 provides a unique mechanism to dynamically optimize both ionic and electronic transport, thereby improving the capacity, rate capability, and overall electrochemical performance under thermal activation.
Flower-like VO2 (B) microspheres self-assembled from ultrathin nanosheets were synthesized by Liang et al. (Figure 3c). The resulting three-dimensional nanostructures exhibited a specific surface area of up to 30.05 m2 g−1, which significantly enhanced the lithium-ion transport kinetics [45]. When used as a cathode material, the initial discharge specific capacity reached 209.6 mAh g−1 at a high current density of 1 A g−1, with a capacity retention of 83.1% after 200 cycles. This excellent performance was attributed to the synergistic effects of the three-dimensional structure in shortening the ion diffusion pathways, buffering volume expansion, and enhancing the electrode–electrolyte interface. In addition, heterostructures were constructed by Jang et al. to effectively integrate the performance advantages of multiple materials. A one-dimensional/two-dimensional (1D/2D) heterostructure composed of VO2 (B) nanowires and g-C3N4 nanosheets [46] was synthesized via a hydrothermal method (Figure 3d). The resulting composite exhibited a high specific capacity of up to 779 mAh g−1 at a current density of 0.1 A g−1, significantly outperforming pure VO2 (601.4 mAh g−1). This enhanced performance was attributed to the high electrical conductivity of g-C3N4, the rapid ion transport provided by VO2 (B) nanowires, and their synergistic effect in improving structural stability, which effectively mitigated damage caused by volumetric changes during cycling.
VO2 has also been extensively investigated as an anode material for lithium-ion batteries; however, the low initial coulombic efficiency (ICE) and unstable solid electrolyte interphase (SEI) during cycling remain major obstacles to its practical application. To address this challenge, a nondestructive chemical prelithiation strategy was proposed by Yan et al. to significantly improve the ICE of VO2 (B) (approaching 100%) and enhance both its cycling stability and overall electrochemical performance [47]. Chemical prelithiation using a lithium aromatic reagent produced a thinner and more uniform SEI film enriched with highly conductive LiF components, which effectively suppressed side reactions (Figure 3f). The prelithiated VO2 (B) electrode retained a high reversible capacity of 375 mAh g−1 after 1000 cycles at a current density of 1.0 A g−1. The application of a VO2 (D) submicron spherical hierarchical structure as an anode material for aqueous lithium-ion batteries was first reported by Ma et al. [48]. The material was synthesized via a template-free solvothermal method and demonstrated good structural stability and electrochemical performance. The constructed VO2 (D)/LiMn2O4 full cell delivered a high specific capacity of up to 97.43 mAh g−1 across a wide voltage window (0.2–1.8 V). Mechanistic analysis indicated that the unique three-dimensional structure of VO2 (D) facilitates ion and electron transport, which expands the application prospects of polycrystalline VO2 (D) in energy storage systems (Figure 3e).
VO2-based electrodes are prone to failure because of structural degradation during cycling. Repeated ion insertion and extraction cause lattice strain and microcracks, compromising the reversibility of the material. A NaV6O15@VO2 (M)@V2C composite with a hierarchical two-dimensional architecture was developed by Tan et al., in which VO2 (M) nanosheets are surrounded by NaV6O15 nanorods and embedded in a dual-conductive network of V2C MXenes [49], effectively enhancing the lithium-ion storage capacity and structural stability (Figure 3g). When applied as an anode in lithium-ion batteries, the material achieved a reversible capacity of 408.1 mAh g−1 after 100 cycles at a current density of 100 mA g−1 and retained 204.5 mAh g−1 after 400 cycles at 1 A g−1, with a coulombic efficiency of 99.63%. The conductive network not only accelerates lithium-ion migration but also plays a crucial role in mitigating structural collapse by effectively alleviating the stress induced by volume changes during cycling. This modification reinforces structural stability, thereby enhancing the durability and performance of the electrode.
Figure 3. Simulated crystal structure evolution during Li insertion: (a) VO2 (M); (b) VO2 (R). Reproduced with permission [44]. Copyright 2024, Royal Society of Chemistry. (c) SEM images of flower-like VO2 (B) microspheres. Reproduced with permission [45]. Copyright 2024, Elsevier. (d) GCD profiles of the VO2, g-C3N4, and composite electrodes at a 0.1 C current density. Reproduced with permission [46]. Copyright 2023, Elsevier. (e) Cycling performance of the VO2 (D)/LiMn2O4 full cell at 100 and 200 mA g−1. Reproduced with permission [48]. Copyright 2022, Elsevier. (f) Schematic of the VO2 (B) prelithiation process and its structural advantages over pristine VO2 (B). Reproduced with permission [47]. Copyright 2024, Elsevier. (g) Schematic illustration of the preparation of the NaV6O15@VO2@V2C composite. Reproduced with permission [49]. Copyright 2023, Elsevier.
Figure 3. Simulated crystal structure evolution during Li insertion: (a) VO2 (M); (b) VO2 (R). Reproduced with permission [44]. Copyright 2024, Royal Society of Chemistry. (c) SEM images of flower-like VO2 (B) microspheres. Reproduced with permission [45]. Copyright 2024, Elsevier. (d) GCD profiles of the VO2, g-C3N4, and composite electrodes at a 0.1 C current density. Reproduced with permission [46]. Copyright 2023, Elsevier. (e) Cycling performance of the VO2 (D)/LiMn2O4 full cell at 100 and 200 mA g−1. Reproduced with permission [48]. Copyright 2022, Elsevier. (f) Schematic of the VO2 (B) prelithiation process and its structural advantages over pristine VO2 (B). Reproduced with permission [47]. Copyright 2024, Elsevier. (g) Schematic illustration of the preparation of the NaV6O15@VO2@V2C composite. Reproduced with permission [49]. Copyright 2023, Elsevier.
Nanomaterials 15 01167 g003

3.2. Zinc-Ion Batteries

Zinc-ion batteries are among several battery systems that are adaptable to both aqueous and nonaqueous electrolytes. Compared with other energy storage technologies, these systems have garnered significant attention because of their high safety, environmental friendliness, and low cost. Unlike lithium-ion systems, ZIBs operate in neutral or mildly acidic aqueous electrolytes, which present new challenges for electrode materials, particularly with respect to structural stability, Zn2+ diffusion kinetics, and redox reversibility. The unique layered and tunneled crystal structure, multivalent transition capability, and rapid ion transport characteristics of VO2 make it a promising candidate for Zn2+ storage.
He et al. modulated the material morphology to optimize the tunneling orientation in electrodes, and (00l) facet-dominated VO2 (B) nanobelts with dispersion (VO2-D) were fabricated. These electrodes exhibited excellent rate performance and cycling stability due to their c-axis-oriented tunneling structure (Figure 4a). A specific capacity of 420.8 mAh g−1 at 0.1 A g−1 and 344.8 mAh g−1 at 10 A g−1 was achieved, with a capacity retention of 84.3% after 5000 cycles. This offers a novel strategy to enhance ion transport kinetics in tunneling vanadium oxides through the concurrent modulation of exposed crystal facets and morphology-dependent electrode architectures [50]. Pinnock et al. enhanced the performance of aqueous ZIBs by increasing the specific capacity of VO2 (B) from 310 to 500 mAh g−1 through the optimization of hydrothermal synthesis and freeze–drying treatment. The optimized cathode achieved a stable capacity retention of 71.5% after 1000 cycles [51]. Furthermore, ultrahigh-loading three-dimensional electrodes with a mass loading of up to 24 mg cm−2 were fabricated by depositing VO2 (B) onto porous glassy carbon foam (Figure 4b), achieving an area capacitance of 4.15 mAh cm−2 at 1 mA cm−2 and maintaining 81.5% capacity retention after 1000 cycles. These improvements were attributed to the large surface area and excellent ion permeability provided by the 3D structure, which effectively alleviated performance degradation under high-loading conditions. Yeon et al. developed a binder-free VO2 composite electrode utilizing polydopamine-derived pyrolytic protein fibers (pp-fibers) as a flexible current collector, enabling enhanced flexibility and electrochemical stability [52]. The binder and additive-free system was synthesized via the hydrothermal growth of VO2 (B) nanosheets on pp-fibers. The electrode demonstrated outstanding electrochemical properties in aqueous ZIBs, delivering a specific capacity of 491 mAh g−1 at 0.2 A g−1 (Figure 4c) and a minimal capacity decay rate of 0.001% per cycle over 20,000 cycles at 1 A g−1. Moreover, the assembled flexible pouch cells remained operational under mechanical deformation, offering a promising strategy for the development of flexible energy storage devices.
Vanadium vacancies were introduced into tunneled VO2 via hydrothermal synthesis and chemical etching, enabling the modulation of the vanadium valence state and lattice contraction to improve structural stability [53]. These vacancies induce the formation of high-valence vanadium ions, alter the surface charge distribution, and introduce abundant electrochemically active sites that increase the capacity. Furthermore, ion diffusion and electron transport are facilitated through the reduction of the Zn2+ migration barrier and charge transfer activation energy. The modified VO2 cathode maintained a capacity of 332 mAh g−1 after 200 cycles at 0.1 A g−1, demonstrating superior cycling stability at low current densities (Figure 4d).
Although oxygen vacancies (Ov) can expand the lattice and promote Zn2+ intercalation, they are prone to being filled by oxygen from the electrolyte and may contribute to capacity degradation due to localized electric field migration. While heterostructure interfaces have shown the ability to modulate the electronic structure of Ovs, effective strategies to improve their stability remain underdeveloped. Fang et al. constructed covalent heterostructures featuring Ti-O-V asymmetric orbital hybridization by growing VO2 nanowalls on MXene nanosheets via a H2O2-assisted hydrothermal process (MXene-VO2−x) [54]. This interfacial orbital hybridization promoted electron transfer from VO2 to the MXene, thereby stabilizing the oxygen vacancies both thermodynamically and kinetically (Figure 4e). A reversible capacity of 487.9 mAh g−1 at 0.2 A g−1 and a retention rate of 98.6% after 30,000 cycles at 20 A g−1 were achieved. Moreover, the flexible devices remained functional under mechanical deformation, offering a novel orbital engineering approach for the rational design of highly reversible ZIB cathodes.
Current strategies to increase the performance of VO2-based zinc-ion batteries involve multidimensional approaches, including elemental doping, structural modulation, and heterostructure construction. Table 2 summarizes recent VO2-based cathode materials along with their test conditions and electrochemical performance metrics, serving as a valuable reference for ongoing and future research in this field.
As shown in Table 2, VO2-based cathodes clearly exhibit trade-offs between capacity, stability, and structural complexity. High-performance composites such as VO2/V2C@CNF and MXene-VO2−x deliver outstanding specific capacities (up to ~550 mAh g−1) and excellent cycling stability (over 85% retention for thousands of cycles) [54,55], primarily due to enhanced electronic conductivity and reinforced structural integrity from carbonaceous frameworks. However, these systems often involve complex synthesis processes, which may hinder large-scale application. In contrast, simpler VO2 phases such as VO2 (B) offer more accessible fabrication routes but typically suffer from lower long-term stability and capacity fading. Surface functionalization strategies (e.g., H-VO2@GO) effectively buffer volume changes and improve ion transport, resulting in significantly enhanced capacity retention [56]. Defect engineering, as in Ov-CoVO, also contributes to improved reaction kinetics and exceptional stability (~97% retention) [57]. Moreover, some materials emphasize specific performance metrics at the expense of others. For example, D-VO2 has a high initial capacity but poor retention at high rates [53], whereas VO@NDA sacrifices the capacity to achieve an ultralong cycling life [58].
These examples collectively underscore that while various modification strategies can greatly enhance VO2 cathode performance, they often entail trade-offs among energy density, rate capability, structural stability, and scalability. Therefore, a rational design must balance these factors to meet the demands of practical zinc-ion battery applications.
Table 2. Recent progress in the use of VO2 and its modified materials as cathode materials for zinc-ion batteries.
Table 2. Recent progress in the use of VO2 and its modified materials as cathode materials for zinc-ion batteries.
MaterialsCell TypeVoltage Range (V)Capacity
(mA h g−1)
Capacity Retention
(Cycles)
Ref.
VO2-Dhalf-cell0.2–1.5420.8 (0.1 A g−1)84.3 % (5 A g−1, 5000)[50]
VO2 (B)half-cell0.2–1.6447 (0.2 A g−1)71.3 % (2 A g−1, 1000)[51]
D-VO2half-cell0.2–1.6332 (0.1 A g−1)46.2 % (20 A g−1, 1800)[53]
VO2half-cell0.3–1.3317 (0.5 A g−1)81.0 % (10 A g−1, 2000)[59]
H-VO2@GOhalf-cell0.2–1.5400.1 (0.5 A g−1)96.1 % (10 A g−1, 1500)[56]
pp-fibers@VO2 (B)half-cell0.2–1.8491 (0.2 A g−1)80.17 % (1 A g−1, 20,000)[52]
CrVOhalf-cell0.2–1.3312.8 (0.1 A g−1)90.39 % (57 A g−1, 2000)[60]
Ov-CoVOhalf-cell0.3–1.4475 (0.2 A g−1)97.5 % (5 A g−1, 3000)[57]
MnVOhalf-cell0.3–1.4209.6 (0.1 A g−1)80.7 % (5 A g−1, 10,000)[61]
Mg-VO2half-cell0.2–1.5385.7 (0.1 A g−1)70.5 % (2 A g−1, 800)[62]
Ce-VO2half-cell0.2–1.4371.4 (0.1 A g−1)85.0 % (5 A g−1, 2000)[63]
Ov-VO2@CNFhalf-cell0.2–1.4450 (0.1 A g−1)85.0 % (5 A g−1, 2000)[64]
V2O3/VO2@S/N-Chalf-cell0.2–1.6257.8 (1 A g−1)81.8 % (200 A g−1, 150,000)[65]
V6O13/VO2half-cell0.2–1.6498.3 (0.2 A g−1)96.8 % (10 A g−1, 5000)[66]
VO2/V2C@CNFhalf-cell0.2–1.7549 (0.1 A g−1)87.0 % (10 A g−1, 5000)[55]
MXene-VO2−xhalf-cell0.2–1.6487.9 (0.2 A g−1)98.6 % (20 A g−1, 30,000)[54]
Mo-VO2half-cell0.4–1.5409.3 (0.1 A g−1)89.5 % (2 A g−1, 1000)[67]
VO@NDAhalf-cell0.4–1.6241 (10 A g−1)97.45 % (10 A g−1, 15,000)[58]
Note. Abbreviations used: “CrVO” refers to Cr-ion-doped VO2 (B); “Ov-CoVO” represents oxygen-deficient Cosubstituted VO2; and “MnVO” indicates Mn-ion-doped VO2.
Figure 4. (a) Schematic illustration of the effect of the material morphology-related electrode arrangement on the tunnel orientation and ion diffusion behavior. Reproduced with permission [50]. Copyright 2020, Wiley-VCH GmbH. (b) GCD curve comparison of 3D and conventional cathodes at 1 mA cm−2. Reproduced with permission [51]. Copyright 2025, Royal Society of Chemistry. (c) GCD curves of pp-fibers@VO2 (B) at current densities from 0.2 to 10.0 A g−1. Reproduced with permission [52]. Copyright 2024, Wiley-VCH GmbH. (d) Long-term cycling of the D-VO2 cathode. Reproduced with permission [53]. Copyright 2025, Elsevier. (e) Schematic diagram of the predicted formation mechanism of the heterointerface. Reproduced with permission [54]. Copyright 2025, Royal Society of Chemistry.
Figure 4. (a) Schematic illustration of the effect of the material morphology-related electrode arrangement on the tunnel orientation and ion diffusion behavior. Reproduced with permission [50]. Copyright 2020, Wiley-VCH GmbH. (b) GCD curve comparison of 3D and conventional cathodes at 1 mA cm−2. Reproduced with permission [51]. Copyright 2025, Royal Society of Chemistry. (c) GCD curves of pp-fibers@VO2 (B) at current densities from 0.2 to 10.0 A g−1. Reproduced with permission [52]. Copyright 2024, Wiley-VCH GmbH. (d) Long-term cycling of the D-VO2 cathode. Reproduced with permission [53]. Copyright 2025, Elsevier. (e) Schematic diagram of the predicted formation mechanism of the heterointerface. Reproduced with permission [54]. Copyright 2025, Royal Society of Chemistry.
Nanomaterials 15 01167 g004

3.3. Photoassisted Batteries

Although solar energy is one of the cleanest and most abundant energy sources, its development and utilization still require significant improvement. Photoassisted batteries, as an emerging technological system that integrates light energy conversion with electrochemical energy storage, exhibit distinct technical advantages through the synergistic optimization of light excitation and battery performance. Its core mechanism involves the use of semiconductors, quantum dots, and other photoactive materials to generate photogenerated carriers under illumination, which promote electrochemical reactions at both electrodes through interfacial charge transfer or directly convert light energy into chemical energy for storage.
Previous studies have demonstrated that VO2 has notable electrochemical advantages in lithium and zinc battery systems, with its multiple valence states and reversible phase transitions offering an ideal platform for ion storage. Owing to its dual functionalities in ion storage and light absorption, the coupling of the semiconductor characteristics of VO2 with its ion intercalation properties presents significant potential for research on light-induced charge injection and storage mechanisms. Extensive research has been conducted on the application of VO2 in photoassisted batteries [68,69], which not only broadens its functional scope but also offers novel theoretical insights for the development of photoelectrochemical synergistic multienergy storage systems.
Recently, numerous innovative studies in the field of photoassisted batteries have emerged. Ding et al. fabricated a highly ordered, vertically aligned C@VO2/ZnO microrod array by combining photolithography with a hydrothermal method to synthesize precursor ZMRAs [70], which were then composited with carbon-coated VO2 to construct a three-dimensional heterojunction network (Figure 5a). The heterostructure facilitates the separation of photogenerated carriers, while the carbon coating improves the electrical conductivity, and the 3D network increases the specific surface area, providing additional Zn2+ storage sites. This synergistic interaction among components resulted in an 18.6% increase in battery capacity under light illumination. The resulting photorechargeable zinc-ion battery was capable of direct light charging without the need for external photovoltaic modules, achieving a photoconversion efficiency of 3.3%. It delivered a capacity of 286.0 mAh g−1 in the dark at a current density of 500 mA g−1, which increased to 339.3 mAh g−1 under illumination, with a capacity retention rate of 88.79% after 300 cycles at 1000 mA g−1. Roy et al. were the first to develop an air-assisted, self-charging energy storage device by employing WO3 as a charge-separating layer in combination with VO2 to form a heterostructure (Figure 5b). In this system, VO2/WO3 serves as the cathode, where the cubic WO3 structure (200–300 nm) is integrated with the micrometer-scale lamellar VO2, thereby increasing the specific surface area of the electrode and optimizing the ion diffusion pathways. Additionally, the high electrical conductivity of WO3 reduces the charge-transfer resistance and enhances the electrochemical activity. The built-in electric field at the heterojunction interface significantly enhances the separation of photogenerated charges [71]. The battery capacity increases by 170% under light exposure at a current density of 0.02 mA cm−2, and the open-circuit voltage reaches 0.9 V within 140 s during air-assisted self-charging (Figure 5c). Yang et al. synthesized V2O5/VO2 hollow nanorods as cathode materials for photocharged zinc-ion batteries via controlled oxidation [72]. In this system, uniquely structured WO3/VO3 hollow nanorods were also employed as cathode materials. The unique heterojunction architecture and hollow morphology enable multiple synergistic effects. The built-in electric field of the type-II heterojunction accelerates the transfer of photogenerated electrons from VO2 to V2O5, whereas the holes remaining in VO2 facilitate Zn2+ deintercalation. The hollow structure enhances light absorption by increasing the specific surface area, resulting in a surface photovoltage of 2573 mV. This design achieved a light-specific capacity of 785.6 mAh g−1 at a current density of 200 mA g−1 (Figure 5d), representing a 77.8% increase compared with the dark state. A photoconversion efficiency of 4.3% was recorded, with 53.1% capacity retention after 4000 cycles at 1000 mA g−1, demonstrating excellent long-term stability.
Although photoassisted batteries currently exhibit relatively low photoconversion efficiencies and are unable to match the performance of conventional PV systems, they present a highly promising avenue for future energy technologies. Their integrated nature offers unique advantages in terms of system simplification and space efficiency. Continued efforts in material innovation and system-level optimization are essential to unlock their full potential and bring them closer to practical deployment.

3.4. Other Batteries and Supercapacitors

The potential of VO2 extends far beyond current applications, as it has also exhibited a promising performance in emerging energy storage systems such as calcium-ion batteries (CIBs), lithium–sulfur batteries, aluminum-ion batteries (AIBs), and supercapacitors. Its polycrystalline structure, reversible multivalent redox behavior, wide electrochemical window, and excellent structural stability make VO2 suitable for diverse ion storage mechanisms. For example, its layered or tunneled structure enables the reversible intercalation of large-radius ions such as Ca2+ and Al3+; in Li-S batteries, it serves as a polar host to suppress the shuttle effect of lithium polysulfides (LiPS); and in supercapacitors, it delivers outstanding pseudocapacitive performance.
CIBs have garnered interest because of their low redox potential, abundant calcium resources, and low propensity for dendrite formation. Wang et al. fabricated a VO2 (B)/rGO heterojunction cathode in which VO2 (B) offered large tunnels for Ca2+ diffusion [73], while the V-O-C bonding between VO2 (B) and rGO formed a conductive network (Figure 6a), enhancing the electronic conductivity and reducing the charge transfer resistance. The interfacial Ca2+ diffusion barrier was only 0.64 eV (Figure 6b), and the discharge capacity improved from 157.2 to 319.2 mAh g−1. Even after 3000 cycles at 50 °C, 85% of the capacity was retained, demonstrating excellent thermal stability. In lithium–sulfur batteries, VO2 has also been proven capable of regulating LiPS intermediate behavior. Pang et al. synthesized atomically dispersed Fe3+-doped VO2 nanoribbons (Fe-VO2) via a hydrothermal method [74]. Through electronic metal–support interactions (EMSIs), the electronic structure was modulated, increasing the VO2 conductivity by approximately three orders of magnitude. This material significantly reduced the decomposition barrier of Li2S from 1.60 to 1.32 eV and the Li+ diffusion barrier from 1.42 to 0.99 eV (Figure 6c), enabling strong LiPS adsorption and rapid conversion. The material achieved an initial capacity of 1275 mAh g−1 at 0.1 C and maintained 67% of its capacity after 500 cycles at 1 C, offering a new EMSI-based design approach for Li-S battery cathysts (Figure 6d). For AIBs, Wang et al. fabricated Cu-doped VO2 nanoflowers via a glucose-assisted hydrothermal method [75]. Cu2+ doping induced hybridization between the 3d orbitals and V-O orbitals, which enhanced electron coupling and narrowed the band gap from 1.18 eV to nearly zero (Figure 6e,f), improving the electronic conductivity. The material demonstrated an optimized Al3+ adsorption energy of −1.23 eV and a diffusion coefficient of 10–11 cm2 s−1. By tuning the oxygen vacancies, its pseudocapacitive behavior was enhanced. The Cu-VO2 electrode delivered an initial capacity of 642 mAh g−1 at 0.4 A g−1 and maintained 116 mAh g−1 after 200 cycles, exhibiting excellent full-cell performance when paired with a 5 M Al(OTF)3 electrolyte and an ionic liquid anode.
VO2 has also shown broad applicability in supercapacitors. Chen et al. introduced a Co2+ preinsertion strategy to stabilize the VO2 tunnel structure, significantly inhibiting vanadium dissolution and enhancing both electronic and structural stability [76]. The NH4+ storage mechanism was found to involve reversible intercalation and hydrogen bonding. The resulting Co-VO2 electrode achieved an areal capacitance of 9.5 F cm−2 at 1 mA cm−2 (Figure 6g) and retained 77.4% capacity after 2000 cycles, outperforming pristine VO2. A hybrid device based on Co-VO2/CuFe-PBA achieved an areal capacitance of 3035.8 mF cm−2 with stable cycling (Figure 6h). VO2/V2C MXene composites were further engineered by Zhu et al., who reported enhanced electrical conductivity and a high density of redox-active sites as a result of synergistic interfacial effects [77]. An asymmetric supercapacitor based on this composite delivered an energy density of 10.56 Wh L−1 at a power density of 127.8 W L−1 (Figure 6i), with a capacity retention of 74.2% after 5000 cycles, offering a novel strategy for the development of high-performance electrodes.
Overall, ongoing research on VO2 across various energy storage technologies continues to expand its material boundaries and provides critical insights and material support for designing high-performance and highly stable next-generation energy storage systems.

4. Summary and Outlook

Owing to its unique semiconductor properties, temperature-sensitive phase transition behavior, and highly tunable crystal structure, VO2 offers a crucial material foundation for the development of high-performance energy storage devices and demonstrates multidimensional application potential. In this paper, we systematically review the crystal structure of VO2, its mainstream synthesis methods, and the latest research progress in the field of energy storage. Currently, VO2 has been widely applied in lithium-ion batteries, zinc-ion batteries, photoassisted batteries, and supercapacitors. Its conductivity and structural stability have been effectively enhanced through elemental doping and composite modification. However, several challenges remain in its practical applications, including structural degradation during charge and discharge cycles, high costs associated with large-scale production, and an unclear correlation between the phase transition process and energy storage mechanisms.
In the future, the development of VO2 in energy storage systems may achieve breakthroughs in the following aspects:
  • Material design:
Conventional trial-and-error methods for tuning VO2 crystal structures are often inefficient and costly, making them inadequate for meeting the increasing demand for rapid performance optimization in energy storage applications. In contrast, a machine learning-driven design strategy integrated with high-throughput computational and experimental datasets offers a way to analyze the complex correlations between structural parameters and electrochemical performance in detail. Such theoretical models enable the rapid screening of critical factors, including dopant species, concentrations, and growth conditions, thereby facilitating the efficient optimization of material design. Through iterative model refinement, this data-driven approach can continuously generate novel VO2 structural configurations, significantly reduce the research and development cycle, and accelerate the paradigm shift from empirical to predictive material design.
2.
Electrode construction:
For electrode structure engineering, nanoscale array architectures can be designed to optimize the exposure of active crystal facets and improve the utilization of electrochemical reaction sites; alternatively, 3D porous electrode structures can be fabricated to increase interfacial contact and promote ion transport efficiency, as exemplified by the design of 3D bionanostructures, which represents a promising direction for significant innovation.
3.
Research methodology:
Characterization technologies are essential tools for investigating energy storage materials, as they reveal reaction mechanisms and structural evolution. By employing advanced in situ and synchrotron radiation-based techniques, it is possible to dynamically track phase transitions, morphological changes, and interfacial reactions during the energy storage process, thereby enabling a deeper understanding of the intrinsic relationship between structure and performance. For example, in situ X-ray diffraction can monitor real-time crystal phase transitions of electrode materials during charge–discharge cycles; in situ transmission electron microscopy enables a direct visualization of structural and morphological evolution at the nanoscale; and in situ Raman spectroscopy can track chemical bond changes associated with phase transformations. Additionally, synchrotron-based X-ray absorption spectroscopy, including both the X-ray absorption near-edge structure and extended X-ray absorption fine structure, offers precise insights into the valence states and local coordination environments of transition metals, providing critical support for understanding the electronic structure evolution of materials under dynamic electrochemical conditions. The synergistic application of these techniques facilitates a comprehensive elucidation of the coupling mechanisms between structural changes and electrochemical performance, offering a robust theoretical and data-driven foundation for the compositional design and process optimization of energy storage materials.
4.
Industrialization:
The development of low-cost, environmentally friendly, scalable, and robust synthesis methods is essential for advancing the practical application of VO2 materials in secure and intelligent energy storage systems. Such methods ensure consistent material quality during large-scale production, while optimizing synthesis routes and adopting green technologies can further reduce costs and increase the commercialization potential of VO2-based energy storage devices.

Author Contributions

Y.H.: Writing—review and editing, Writing—original draft, Conceptualization, X.G.: Writing—review and editing, Conceptualization, J.L.: Writing—review and editing, Conceptualization, J.Z.: Conceptualization, J.W.: Conceptualization, D.L.: Conceptualization, S.Z.: Conceptualization, W.F.: Writing—review and editing, Writing—original draft, Conceptualization. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Heilongjiang Province, China (Grant No. PL2024E001). And The APC was funded by MDPI.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tang, L.; Peng, H.; Kang, J.; Chen, H.; Zhang, M.; Liu, Y.; Kim, D.H.; Liu, Y.; Lin, Z. Zn-Based batteries for sustainable energy storage: Strategies and mechanisms. Chem. Soc. Rev. 2024, 53, 4877–4925. [Google Scholar] [CrossRef]
  2. Huang, J.; Xie, Y.; You, Y.; Yuan, J.; Xu, Q.; Xie, H.; Chen, Y. Rational Design of Electrode Materials for Advanced Supercapacitors: From Lab Research to Commercialization. Adv. Funct. Mater. 2023, 33, 2213095. [Google Scholar] [CrossRef]
  3. Guo, A.; Wang, Z.; Chen, L.; Liu, W.; Zhang, K.; Cao, L.; Liang, B.; Luo, D. A Comprehensive Review of the Mechanism and Modification Strategies of V2O5 Cathodes for Aqueous Zinc-Ion Batteries. ACS Nano 2024, 18, 27261–27286. [Google Scholar] [CrossRef]
  4. Liu, J.; Gao, X.; Zhang, Y.; Zhang, N.; Shao, Z.; Sun, R.; Zhou, J.; Li, D.; Xiang, X.; Feng, W. Boosted energy capacity and cycling stability of photo-assisted zinc-ion Batteries by K+ intercalated hydrated vanadium oxide-based photocathodes. J. Energy Storage 2025, 130, 117461. [Google Scholar] [CrossRef]
  5. Zhanadilov, O.; Kim, H.J.; Konarov, A.; Jeong, J.; Park, J.-H.; Chung, K.Y.; Bakenov, Z.; Yashiro, H.; Myung, S.-T. Layered manganese oxide cathode boosting high-capacity and long-term cyclability in aqueous Zinc-Ion batteries. Energy Storage Mater. 2024, 67, 103283. [Google Scholar] [CrossRef]
  6. Gao, X.; Tian, D.; Shi, Z.; Zhang, N.; Sun, R.; Liu, J.; Tsai, H.; Xiang, X.; Feng, W. An Efficient MnO2 Photocathode with an Excellent SnO2 Electron Transport Layer for Photo-accelerated Zinc Ion Batteries. Small 2024, 20, 2405627. [Google Scholar] [CrossRef]
  7. Tian, D.; Gao, X.; Wang, J.; Sun, R.; Liu, J.; Cao, J.; Feng, W. An optimized MnO2 photocathode by doping engineering for high capacity and stability photo-assisted zinc-ion batteries. J. Mater. Chem. A 2025, 13, 22057–22065. [Google Scholar] [CrossRef]
  8. Yi, P.; Li, Z.; Ma, L.; Feng, B.; Liu, Z.; Liu, Y.; Lu, W.; Cao, S.; Fang, H.; Ye, M.; et al. Eco-friendly High-performance Symmetric all-COF/Graphene Aqueous Zinc-ion Batteries. Adv. Mater. 2024, 36, 2414379. [Google Scholar] [CrossRef]
  9. Cho, B.; Huh, S.; Kim, S.H.; Yu, S.; Bae, J.; Yoo, J.; Yu, S. Long cycle-life aqueous Zn battery enabled by facile carbon nanotube coating on Cu current collector. Carbon Energy 2024, 6, e441. [Google Scholar] [CrossRef]
  10. Yuan, W.; Weng, J.; Ding, M.; Jiang, H.-M.; Fan, Z.; Zhao, Z.; Zhang, P.; Xu, L.-P.; Zhou, P. Novel covalent organic framework/carbon nanotube composites with multiple redox-active sites for high-performance Na storage. Energy Storage Mater. 2024, 65, 103142. [Google Scholar] [CrossRef]
  11. Li, Y.X.; Feng, Y.S.; Li, L.X.; Yin, X.; Cao, F.F.; Ye, H. Green synthesis and applications of MXene for lithium–sulfur batteries. Energy Storage Mater. 2024, 67, 103257. [Google Scholar] [CrossRef]
  12. Zhang, G.; Yu, H.; Li, X.; Zhang, X.; Hou, C.; Sun, S.; Du, Y.; Guo, Z.; Dang, F. Construction of MnS/MoS2 heterostructure on two-dimensional MoS2 surface to regulate the reaction pathways for high-performance Li-O2 batteries. J. Energy Chem. 2024, 93, 443–452. [Google Scholar] [CrossRef]
  13. Housel, L.M.; Quilty, C.D.; Abraham, A.; Tang, C.R.; McCarthy, A.H.; Renderos, G.D.; Liu, P.; Takeuchi, E.S.; Marschilok, A.C.; Takeuchi, K.J. Investigation of Conductivity and Ionic Transport of VO2 (M) and VO2 (R) via Electrochemical Study. Chem. Mater. 2018, 30, 7535–7544. [Google Scholar] [CrossRef]
  14. Li, W.; Huang, J.; Cao, L.; Li, X.; Chen, S.; Feng, L. Polycrystalline VO2 (M) with well-dispersed crystalline zones for enhanced electroactivity of Lithium-ion batteries. J. Alloys Compd. 2020, 812, 152122. [Google Scholar] [CrossRef]
  15. Shvets, P.; Shabanov, A.; Maksimova, K.; Goikhman, A. Micro-Raman mapping of VO2 (T) microcrystals orientation. Vib. Spectrosc. 2022, 118, 103328. [Google Scholar] [CrossRef]
  16. Oka, Y.; Sato, S.; Yao, T.; Yamamoto, N. Crystal Structures and Transition Mechanism of VO2 (A). J. Solid State Chem. 1998, 141, 594–598. [Google Scholar] [CrossRef]
  17. Cheng, B.; Zhang, H.; Li, Q.; Liu, J.; Liu, B. Morphology Tuned Pressure Induced Amorphization in VO2 (B) Nanobelts. Inorganics 2022, 10, 122. [Google Scholar] [CrossRef]
  18. Hagrman, D.; Zubieta, J.; Warren, C.J.; Meyer, L.M.; Treacy, M.M.J.; Haushalter, R.C. A New Polymorph of VO2 Prepared by Soft Chemical Methods. J. Solid State Chem. 1998, 138, 178–182. [Google Scholar] [CrossRef]
  19. Qu, B.Y.; Liu, L.; Xie, Y.; Pan, B.C. Theoretical study of the new compound VO2 (D). Phys. Lett. A 2011, 375, 3474–3477. [Google Scholar] [CrossRef]
  20. Wu, C.; Hu, Z.; Wang, W.; Zhang, M.; Yang, J.; Xie, Y. Synthetic paramontroseite VO2 with good aqueous Lithium-ion battery performance. Chem. Commun. 2008, 7, 3891–3893. [Google Scholar] [CrossRef]
  21. Liu, Y.-Y.; Lv, T.-T.; Wang, H.; Guo, X.-T.; Liu, C.-S.; Pang, H. Nsutite-type VO2 microcrystals as highly durable cathode materials for aqueous zinc-Ion batteries. Chem. Eng. J. 2021, 417, 128408. [Google Scholar] [CrossRef]
  22. Khan, Z.; Singh, P.; Ansari, S.A.; Manippady, S.R.; Jaiswal, A.; Saxena, M. VO2 Nanostructures for Batteries and Supercapacitors: A Review. Small 2021, 17, 2006651. [Google Scholar] [CrossRef]
  23. Wan, B.; Wang, Y.; Chen, X.; Zhan, C.; Jiang, H.; Liu, J.-H.; Gao, Y.; Jiang, X.; Cao, X.; Zhang, H.; et al. Critical issues and optimization strategies of vanadium dioxide-based cathodes towardds high-performance aqueous Zn-ion batteries. Chem. Sci. 2025, 16, 8217–8239. [Google Scholar] [CrossRef]
  24. Hou, J.-F. Disorder/order-heterophase VO2 for enhanced lithium storage performance in lithium-ion capacitors. J. Energy Storage 2024, 81, 110433. [Google Scholar] [CrossRef]
  25. Pan, A.; Wu, H.B.; Yu, L.; Lou, X.W. (David). Template-free Synthesis of VO2 Hollow Microspheres with Various Interiors and Their Conversion into V2O5 for Lithium-ion Batteries. Angew. Chem. Int. Ed. 2013, 52, 2226–2230. [Google Scholar] [CrossRef]
  26. Zheng, C.; Zhang, X.; Zhang, J.; Liao, K. Preparation and Characterization of VO2 Nanopowders. J. Solid State Chem. 2001, 156, 274–280. [Google Scholar] [CrossRef]
  27. Jung, D.; Kim, U.; Cho, W. Fabrication of pure monoclinic VO2 nanoporous nanorods via a mild pyrolysis process. Ceram. Int. 2018, 44, 6973–6979. [Google Scholar] [CrossRef]
  28. Lai, X.; Li, Y.; Yang, Y.; Zhou, Z.; Wang, C. Vanadium dioxide (IV) (B) nanostructures prepared via electrodeposition for zinc-Ion storage in electrochemical energy storage devices. J. Energy Storage 2024, 102, 114157. [Google Scholar] [CrossRef]
  29. Xiang, W.; Le Drogoff, B.; Koch, D.; Margot, J.; Chaker, M. High-quality VO2 films synthesized on polymer substrates using room-temperature pulsed laser deposition and annealing. Ceram. Int. 2024, 50, 838–846. [Google Scholar] [CrossRef]
  30. Kim, J.; Haeng Lee, K.; Seo, D.-B.; Jang, H.; Kang, S.; Yim, S.; Sook Lee, S.; An, K.-S. Two-dimensional VO2 nanosheet converted from MXene for flexible thermochromic smart windows. Chem. Eng. J. 2023, 477, 147014. [Google Scholar] [CrossRef]
  31. Huang, Y.; Zhang, D.; Liu, Y.; Jin, J.; Yang, Y.; Chen, T.; Guan, H.; Fan, P.; Lv, W. Phase transition analysis of thermochromic VO2 thin films by temperature-dependent Raman scattering and ellipsometry. Appl. Surf. Sci. 2018, 456, 545–551. [Google Scholar] [CrossRef]
  32. Wu, H.; Yu, Y.; Yu, Y.; Shi, J.; Morikawa, H.; Zhu, C. A facile method for continuous production of temperature-adaptive hyperthermal management core-sheath polyurethane nanofiber yarns based on vanadium dioxide toward commercialization. J. Energy Storage 2024, 86, 111311. [Google Scholar] [CrossRef]
  33. Chen, Z.; Tang, Y.; Ji, A.; Zhang, L.; Gao, Y. Large-Scale Preparation of Durable VO2 Nanocomposite Coatings. ACS Appl. Nano Mater. 2021, 4, 4048–4054. [Google Scholar] [CrossRef]
  34. Man, P.; Zhang, Q.; Sun, J.; Guo, J.; Wang, X.; Zhou, Z.; He, B.; Li, Q.; Xie, L.; Zhao, J.; et al. Hierarchically structured VO2@PPy core-shell nanowire arrays grown on carbon nanotube fibers as advanced cathodes for high-performance wearable asymmetric supercapacitors. Carbon 2018, 139, 21–28. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Zhang, J.; Zhong, Y.; Yu, L.; Deng, Y.; Huang, C.; Liu, X. Direct fabrication of organic carbon coated VO2 (B) (VO2 (B)@C) core-shell structured nanobelts by one step hydrothermal route and its formation mechanism. Appl. Surf. Sci. 2012, 263, 124–131. [Google Scholar] [CrossRef]
  36. Lin, S.; Wang, Y.; Chen, Y.; Cai, Z.; Xiao, J.; Muhmood, T.; Hu, X. Dendrite-Free and Ultra-Long-Life Lithium Metal Anode Enabled via a Three-Dimensional Ordered Porous Nanostructure. ACS Appl. Mater. Interfaces 2021, 13, 41744–41752. [Google Scholar] [CrossRef]
  37. Bai, M.-X.; He, Z.-H.; Hou, J.-F.; Gao, J.-F.; Kong, L.-B. Solvothermally prepared hydrated VO2 (B) for aqueous zinc ion batteries with high capacity and excellent rate capability. J. Alloys Compd. 2023, 936, 168218. [Google Scholar] [CrossRef]
  38. Chen, X.; Wang, L.; Li, H.; Cheng, F.; Chen, J. Porous V2O5 nanofibers as cathode materials for rechargeable aqueous zinc-ion batteries. J. Energy Chem. 2019, 38, 20–25. [Google Scholar] [CrossRef]
  39. Selvam, T.; Dhinasekaran, D.; Subramanian, B.; Rajendran, A.R. Layered Structures of Enriched V5+ States of Vanadium Oxide as a Hybrid Cathode Material for Long-Cyclable Aqueous Zinc-Ion Batteries. ACS Appl. Mater. Interfaces 2023, 15, 30350–30359. [Google Scholar] [CrossRef]
  40. Li, M.; Li, X.; Qin, G.; Luo, K.; Lu, J.; Li, Y.; Liang, G.; Huang, Z.; Zhou, J.; Hultman, L.; et al. Halogenated Ti3C2 MXenes with Electrochemically Active Terminals for High-Performance Zinc Ion Batteries. ACS Nano 2021, 15, 1077–1085. [Google Scholar] [CrossRef]
  41. Shao, F.; Huang, Y.; Wang, X.; Li, Z.; Huang, X.; Huang, W.; Dong, L.; Kang, F.; Liu, W.; Xu, C. MoS2 with high 1T phase content enables fast reversible zinc-ion storage via pseudocapacitance. Chem. Eng. J. 2022, 448, 137688. [Google Scholar] [CrossRef]
  42. Lin, S.; Chen, Y.; Wang, Y.; Cai, Z.; Xiao, J.; Muhmood, T.; Hu, X. Three-Dimensional Ordered Porous Nanostructures for Lithium-Selenium Battery Cathodes That Confer Superior Energy-Storage Performance. ACS Appl. Mater. Interfaces 2021, 13, 9955–9964. [Google Scholar] [CrossRef]
  43. Xiao, J.; Cai, Z.; Muhmood, T.; Hu, X.; Lin, S.; Hu, X. Tailoring Ordered Porous Carbon Embedded with Cu Clusters for High-energy and Long-lasting Phosphorus Anode. Small 2022, 18, 2106930. [Google Scholar] [CrossRef] [PubMed]
  44. Castro-Pardo, S.; Puthirath, A.B.; Fan, S.; Saju, S.; Yang, G.; Nanda, J.; Vajtai, R.; Tang, M.; Ajayan, P.M. VO2 phase change electrodes in Li-ion batteries. J. Mater. Chem. A 2024, 12, 2738–2747. [Google Scholar] [CrossRef]
  45. Liang, F.; Zou, Z.; Zhang, S.; Chen, M.; Yu, F.; Jia, S.; Nong, J. Ultrathin VO2 (B) nanosheets assembled into 3D micro/nano structured flower-like microspheres for high-performance cathode materials of Li-ion batteries. J. Energy Storage 2024, 77, 109787. [Google Scholar] [CrossRef]
  46. Jang, W.Y.; Reddy, C.V.; Wang, R.; Choi, J.; Son, J.; Kakarla, R.R.; Aminabhavi, T.M.; Shim, J. Synthesis of 1D/2D VO2 (B) nanowire/g-C3N4 hybrid architectures as cathode materials for high-performance Li-ion batteries. Chem. Eng. J. 2023, 476, 146786. [Google Scholar] [CrossRef]
  47. Yan, B.; Xu, C.; Liu, L.; Wang, F.; Xiao, W.; Zhang, L.; Yang, X.; Li, X.; Wang, R. NonNon-destructive chemical prelithiation of VO2 (B) anodes constructing a multifunctional interphase layer for advanced Li-Ion batteries. Chem. Eng. J. 2024, 489, 151185. [Google Scholar] [CrossRef]
  48. Ma, Y.; Zhu, D.; Zhou, H.; Tang, Y.; Hu, C.; Meng, X.; Jin, X.; Xu, T.; Cao, X. VO2 (D) submicron-spherical hierarchical structures as novel anode materials for aqueous lithium-ion batteries. J. Alloys Compd. 2023, 930, 167472. [Google Scholar] [CrossRef]
  49. Tan, L.; Wu, J.; Guan, Y.; Jin, Y.; Xu, Z.; Zhu, H.; Zhang, Q.; Li, X.; Dong, Z.; Cong, Y. Dual conductive confinement effects on enhancing Li-ion storage of NaV6O15@VO2 (M)@V2C heterojunction. J. Alloys Compd. 2023, 964, 171242. [Google Scholar] [CrossRef]
  50. He, Q.; Hu, T.; Wu, Q.; Wang, C.; Han, X.; Chen, Z.; Zhu, Y.; Chen, J.; Zhang, Y.; Shi, L.; et al. Tunnel-Oriented VO2 (B) Cathode for High-Rate Aqueous Zinc-Ion Batteries. Adv. Mater. 2024, 36, 2400888. [Google Scholar] [CrossRef]
  51. Pinnock, I.P.; Fan, Y.; Zhu, Y.; Narayan, B.; Wang, T.; Parkin, I.P.; Deka Boruah, B. Advancing high capacity 3D VO2 (B) cathodes for improved zinc-ion battery performance. J. Mater. Chem. A 2025, 13, 1372–1383. [Google Scholar] [CrossRef]
  52. Yeon, J.S.; Park, S.K.; Kim, S.; Mohite, S.V.; Il Kim, W.; Jang, G.; Jang, H.; Bae, J.; Lee, S.M.; Hong, W.G.; et al. Self-supported VO2 on polydopamine-derived pyroprotein-based fibers for ultrastable and flexible aqueous zinc-ion batteries. Carbon Energy 2024, 6, e469. [Google Scholar] [CrossRef]
  53. Liu, X.; Meng, L.; Xu, Z.; Wu, J.; Wang, K.; Zhang, L.; Yu, C.; Li, L. Engineering vanadium vacancies to accelerate ion kinetics for high performance zinc ion battery. J. Colloid Interface Sci. 2025, 684, 439–448. [Google Scholar] [CrossRef]
  54. Fang, Y.; Qi, C.; Bao, W.; Xu, F.; Sun, W.; Liu, B.; Yu, X.; Wang, L.; Jiang, W.; Qiu, P.; et al. Asymmetric orbital hybridization at the MXene-VO2-x interface stabilizes oxygen vacancies for enhanced reversibility in aqueous zinc-ion batteries. Energy Environ. Sci. 2025, 18, 367–377. [Google Scholar] [CrossRef]
  55. Zhang, C.; Wu, Z.-H.; Yang, Z.-Q.; Yu, Y.-X.; Yang, Y. In situ synthesis of VO2 containing high-valent vanadium via surface oxidation of V2C Mxene for robust near-interface reactions in aqueous zinc-ion batteries. Chem. Eng. J. 2025, 511, 162146. [Google Scholar] [CrossRef]
  56. Li, Y.; Liao, X.; Xie, B.; Li, Y.; Zheng, Q.; Lin, D. Hollow VO2 microspheres anchored on graphene as advanced cathodes for aqueous zinc-ion batteries. J. Colloid Interface Sci. 2024, 662, 404–412. [Google Scholar] [CrossRef]
  57. Wang, Z.; Cui, P.; Wang, X.; Chang, M.; Yu, Y.; You, J.; Hu, F.; Wu, Y.; Zhu, K. Co-Substitution Engineering Boosting the Kinetics and Stablity of VO2 for Zn Ion Batteries. Adv. Funct. Mater. 2024, 34, 2407925. [Google Scholar] [CrossRef]
  58. Song, Z.; Zhao, Y.; Wang, H.; Zhou, A.; Jin, X.; Huang, Y.; Li, L.; Wu, F.; Chen, R. Dual mechanism with graded energy storage in long-term aqueous zinc-Ion batteries achieved using a polymer/vanadium dioxide cathode. Energy Environ. Sci. 2024, 17, 6666–6675. [Google Scholar] [CrossRef]
  59. Wang, W.; Feng, C.; Lei, L.; Yang, X.; Li, X.; Ma, L.; Zhang, M.; Fan, H. Regulating Crystal Orientation in VO2 for Aqueous Zinc Batteries with Enhanced Pseudocapacitance. ACS Appl. Mater. Interfaces 2024, 16, 10009–10018. [Google Scholar] [CrossRef] [PubMed]
  60. Chen, X.; Zhai, X.; Wu, Y.; Wang, X.; Zhang, L.; Shang, C.; Zhang, H.; Zhao, C.; Shang, J.; Liu, D. Chromium-doped tunnel-structured VO2 (B) nanorods as high-capacity and stable cathode materials for aqueous zinc-ion batteries. J. Energy Storage 2025, 114, 115826. [Google Scholar] [CrossRef]
  61. Deng, S.; Li, H.; Chen, B.; Xu, Z.; Jiang, Y.; Li, C.; Xiao, W.; Yan, X. High performance of Mn-doped VO2 cathode for aqueous zinc-ion batteries: An insight into Zn2+ storage mechanism. Chem. Eng. J. 2023, 452, 139115. [Google Scholar] [CrossRef]
  62. Wang, N.; Liu, H.; Sun, M.; Ren, X.; Hu, L.; Li, Z.; Yao, X.; Gong, Y.; Jia, C. Achieving Wide-Temperature-Range Sustainable Zinc-Ion Batteries via Magnesium-Doped Cathodes and Gel Electrolytes. ACS Sustain. Chem. Eng. 2024, 12, 3527–3537. [Google Scholar] [CrossRef]
  63. Li, H.; Li, X.; Liu, J.; Ding, N.-W.; Zhang, Q.; Lei, S.; Sydorov, D. Synergistically Electronic and Structural Regulation of VO2 by Ce Doping Toward High-Performance Zinc-Ion Batteries. ACS Sustain. Chem. Eng. 2025, 13, 5091–5100. [Google Scholar] [CrossRef]
  64. Zhang, C.; Wu, Z.-H.; Yang, C.-Q.; Guo, X.-Z.; Yu, Y.-X.; Yang, Y. Rational Regulation of Optimal Oxygen Vacancy Concentrations on VO2 for Superior Aqueous Zinc-Ion Battery Cathodes. ACS Appl. Mater. Interfaces 2024, 16, 40903–40913. [Google Scholar] [CrossRef]
  65. Chen, L.; Chen, Z.; Wu, Z.; Zeng, H.; Chen, Y.; Yang, H.; Qian, W.; Huang, J. V2O3/VO2@S/N-C nanofibers with excellent cycling stability and superior rate capability in aqueous zinc ion batteries. J. Energy Chem. 2024, 99, 83–91. [Google Scholar] [CrossRef]
  66. Liu, Y.; Wu, X. High durable aqueous zinc ion batteries by synergistic effect of V6O13/VO2 electrode materials. J. Energy Chem. 2023, 87, 334–341. [Google Scholar] [CrossRef]
  67. Zhang, D.; Yue, Y.; Yang, C.; Limphirat, W.; Zhang, X.; Qin, J.; Cao, J. Kinetics-Boosted and Dissolution-Suppressed Molybdenum-Doped vanadium dioxide for Long-Life Zinc-Ion batteries. Chem. Eng. J. 2025, 506, 160160. [Google Scholar] [CrossRef]
  68. Deka Boruah, B.; Mathieson, A.; Park, S.K.; Zhang, X.; Wen, B.; Tan, L.; Boies, A.; De Volder, M. Vanadium Dioxide Cathodes for High-Rate Photo-rechargeable Zinc-Ion Batteries. Adv. Energy Mater. 2021, 11, 2100115. [Google Scholar] [CrossRef]
  69. Deka Boruah, B.; De Volder, M. Vanadium dioxide-zinc oxide stacked photocathodes for photo-rechargeable zinc-ion batteries. J. Mater. Chem. A 2021, 9, 23199–23205. [Google Scholar] [CrossRef]
  70. Ding, Z.; Jiang, T.; Zhao, W.; Luo, Z.; San, H.; Li, S.; Li, X.; Zhang, L. Photo-rechargeable zinc-ion battery using highly ordered and vertically oriented C@VO2/ZnO microrod arrays. Energy Storage Mater. 2024, 71, 103646. [Google Scholar] [CrossRef]
  71. Roy, R.; Mahendra, G.; Singh, A.K. Photo-assisted self-chargeable aqueous Zn-ion energy storage device. Chem. Eng. J. 2024, 500, 156870. [Google Scholar] [CrossRef]
  72. Yang, H.; Wu, S.; Badalge, A.M.; Li, X.; Wang, Q.; Zhong, Y.L.; Zhu, T. Controlled oxidation of V2O5/VO2 hollow nanospheres as photocathodes for photo-rechargeable zinc ion batteries with an ultrahigh capacity enhancement. J. Mater. Chem. A 2025, 13, 12216–12225. [Google Scholar] [CrossRef]
  73. Wang, Y.; Wang, J.; Zhang, W.; Chao, F.; Li, J.; Kong, Q.; Qiao, F.; Zhang, L.; Huang, M.; An, Q. Promising VO2 (B)/rGO Heterojunction Cathode for Building High-Capacity and Long-Lifespan Ca-Ion Batteries. Adv. Funct. Mater. 2024, 34, 2314761. [Google Scholar] [CrossRef]
  74. Pang, Y.; Xie, M.; Lu, X.; Wan, Z.; Zhong, Z.; Muhammad, W.; Huan, X.L.; Niu, Y.; Zhang, Z.; Lv, W. Electronic Metal-Support Interactions in Atomically Dispersed Fe(III)-VO2 Nanoribbons for High-performance Lithium-Sulfur Batteries. Adv. Funct. Mater. 2024, 34, 2308849. [Google Scholar] [CrossRef]
  75. Wang, Z.; Gu, H.; Li, Z. Orbital-Modulated Cu-Doped VO2 Nanoflowers via Glucose-Assisted Synthesis: Structural Optimization and Electronic Coupling Engineering for High-Capacity Aqueous Aluminum Ion Batteries. Small 2025, 21, e2503861. [Google Scholar] [CrossRef]
  76. Chen, Q.; Tang, Z.; Li, H.; Liang, W.; Zeng, Y.; Zhang, J.; Hou, G.; Tang, Y. Cobalt Ion-Stabilized VO2 for Aqueous Ammonium Ion Hybrid Supercapacitors. ACS Appl. Mater. Interfaces 2024, 16, 18824–18832. [Google Scholar] [CrossRef]
  77. Zhu, Z.; Xia, Q.; Wang, L.; Hu, Q.; Chang, Y.; Shinde, N.M. In situ grown VO2/V2C MXene and its supercapacitor applications. J. Energy Storage 2024, 88, 111484. [Google Scholar] [CrossRef]
Figure 1. Structure of monoclinic VO2 (B): observed along the (a) a-axis and (b) b-axis. Reproduced with permission [17]. Copyright 2022, MDPI. (c) Structure of VO2 (D) with large and small balls representing V and O ions, respectively. (d) Unit cell of VO2 (D). Reproduced with permission [19]. Copyright 2011, Elsevier. (e,f) Schematic geometries of the crystal structure of monoclinic VO2 (M). Reproduced with permission [14]. Copyright 2020, Elsevier. (g) Crystal structure of VO2 (R) along the tunnel axis. The black and red spheres represent V and O atoms, respectively. (h) Crystal structure of VO2 (R) along the lateral axis. Reproduced with permission [13]. Copyright 2018, American Chemical Society.
Figure 1. Structure of monoclinic VO2 (B): observed along the (a) a-axis and (b) b-axis. Reproduced with permission [17]. Copyright 2022, MDPI. (c) Structure of VO2 (D) with large and small balls representing V and O ions, respectively. (d) Unit cell of VO2 (D). Reproduced with permission [19]. Copyright 2011, Elsevier. (e,f) Schematic geometries of the crystal structure of monoclinic VO2 (M). Reproduced with permission [14]. Copyright 2020, Elsevier. (g) Crystal structure of VO2 (R) along the tunnel axis. The black and red spheres represent V and O atoms, respectively. (h) Crystal structure of VO2 (R) along the lateral axis. Reproduced with permission [13]. Copyright 2018, American Chemical Society.
Nanomaterials 15 01167 g001
Figure 5. (a) Schematic of the preparation process for vertically ordered ZMRAs. Reproduced with permission [70]. Copyright 2024, Elsevier. (b) Demonstration of self-rechargeable energy storage using sustainable sources such as solar and air. (c) GCD curves of VO2/WO3 under both dark and light conditions. Reproduced with permission [71]. Copyright 2024, Elsevier. (d) Discharge curves of ZIBs with V2O5/VO2 photocathodes under dark and light conditions at 1000 mA g−1. Reproduced with permission [72]. Copyright 2025, Royal Society of Chemistry.
Figure 5. (a) Schematic of the preparation process for vertically ordered ZMRAs. Reproduced with permission [70]. Copyright 2024, Elsevier. (b) Demonstration of self-rechargeable energy storage using sustainable sources such as solar and air. (c) GCD curves of VO2/WO3 under both dark and light conditions. Reproduced with permission [71]. Copyright 2024, Elsevier. (d) Discharge curves of ZIBs with V2O5/VO2 photocathodes under dark and light conditions at 1000 mA g−1. Reproduced with permission [72]. Copyright 2025, Royal Society of Chemistry.
Nanomaterials 15 01167 g005
Figure 6. (a) Diffusion path distribution of Ca2+ in VO2 (B) along the b direction; (b) distribution of the corresponding diffusion energy barriers of Ca2+ at the VO2 (B)/rGO interface along the a direction. Reproduced with permission [73]. Copyright 2024, Wiley-VCH GmbH. (c) Li2S decomposition energy barriers on the CB, VO2, and Fe-VO2 surfaces. (d) Rate performance of CB, VO2@CB, and Fe-VO2@CB electrodes from 0.1 to 4 C. Reproduced with permission [74]. Copyright 2023, Wiley-VCH GmbH. DOS and PDOS of (e) VO2 and (f) Cu1mmol-VO2. Reproduced with permission [75]. Copyright 2025, Wiley-VCH GmbH. (g) Area capacitance of VO2 and Co-VO2 anodes at 1–20 mA cm−2. (h) GCD curves of A-HSC. Reproduced with permission [76]. Copyright 2024, American Chemical Society. (i) Ragone plot of a VO2/V2C MXene-based asymmetric supercapacitor. Reproduced with permission [77]. Copyright 2024, Elsevier.
Figure 6. (a) Diffusion path distribution of Ca2+ in VO2 (B) along the b direction; (b) distribution of the corresponding diffusion energy barriers of Ca2+ at the VO2 (B)/rGO interface along the a direction. Reproduced with permission [73]. Copyright 2024, Wiley-VCH GmbH. (c) Li2S decomposition energy barriers on the CB, VO2, and Fe-VO2 surfaces. (d) Rate performance of CB, VO2@CB, and Fe-VO2@CB electrodes from 0.1 to 4 C. Reproduced with permission [74]. Copyright 2023, Wiley-VCH GmbH. DOS and PDOS of (e) VO2 and (f) Cu1mmol-VO2. Reproduced with permission [75]. Copyright 2025, Wiley-VCH GmbH. (g) Area capacitance of VO2 and Co-VO2 anodes at 1–20 mA cm−2. (h) GCD curves of A-HSC. Reproduced with permission [76]. Copyright 2024, American Chemical Society. (i) Ragone plot of a VO2/V2C MXene-based asymmetric supercapacitor. Reproduced with permission [77]. Copyright 2024, Elsevier.
Nanomaterials 15 01167 g006
Table 1. Comparative performance of VO2 and representative electrode materials.
Table 1. Comparative performance of VO2 and representative electrode materials.
MaterialsCostCapacity
(mA h g−1)
Stability
(Cycles)
ScalabilityTypical ApplicationsRef.
VO2Moderate400.2 (0.5 A g−1)84.3% (5 A g−1, 6000)ModerateZn-ion, Li-ion, photoassisted batteries[37]
V2O5Low319 (0.02 A g−1)81% (2 A g−1, 500)HighZn-ion, Li-ion, photoassisted batteries[38]
V6O13Moderate394.2 (0.1 A g−1)94% (2 A g−1, 100)ModerateZn-ion, Li-ion batteries[39]
Ti3C2I2High181 (0.25 A g−1)80% (4 A g−1, 700)LowSupercapacitors, Li-ion, Na-ion batteries[40]
MoS2Moderate156 (0.1 A g−1)97.3% (1 A g−1, 500)ModerateLi-ion, Na-ion, Zn-ion batteries[41]
Note. Since VO2 has been more extensively studied in the context of zinc-ion batteries, the capacity values listed in this table refer to its performance in zinc-ion battery systems for consistency in comparison.
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

He, Y.; Gao, X.; Liu, J.; Zhou, J.; Wang, J.; Li, D.; Zhao, S.; Feng, W. Recent Advances in the Application of VO2 for Electrochemical Energy Storage. Nanomaterials 2025, 15, 1167. https://doi.org/10.3390/nano15151167

AMA Style

He Y, Gao X, Liu J, Zhou J, Wang J, Li D, Zhao S, Feng W. Recent Advances in the Application of VO2 for Electrochemical Energy Storage. Nanomaterials. 2025; 15(15):1167. https://doi.org/10.3390/nano15151167

Chicago/Turabian Style

He, Yuxin, Xinyu Gao, Jiaming Liu, Junxin Zhou, Jiayu Wang, Dan Li, Sha Zhao, and Wei Feng. 2025. "Recent Advances in the Application of VO2 for Electrochemical Energy Storage" Nanomaterials 15, no. 15: 1167. https://doi.org/10.3390/nano15151167

APA Style

He, Y., Gao, X., Liu, J., Zhou, J., Wang, J., Li, D., Zhao, S., & Feng, W. (2025). Recent Advances in the Application of VO2 for Electrochemical Energy Storage. Nanomaterials, 15(15), 1167. https://doi.org/10.3390/nano15151167

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