Download PDF
Review  |  Open Access  |  5 Jun 2025

Recent research progress in the application of Na3V2(PO4)2F3 cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization

Views: 67 |  Downloads: 2 |  Cited:  0
Energy Mater. 2025, 5, 500115.
10.20517/energymater.2024.306 |  © The Author(s) 2025.
Author Information
Article Notes
Cite This Article

Abstract

Sodium vanadium fluorophosphate [Na3V2(PO4)2F3, NVPF], as a sodium (Na) super ionic conductor material, has attracted significant interest as a very promising cathode material for sodium-ion batteries due to its structural stability, rapid ion transport capability, and high operating potential. However, the insulating [PO4] units in NVPF isolate the V atoms, which results in low intrinsic electron conductivity and poor overall electrochemical performance, and the high cost also represents critical challenge that has impeded its widespread use. In recent years, the research focus has shifted to an enhancement of scalable fabrication technologies and improvement of operational robustness under extreme conditions. This has meant a realignment of the research paradigm from the modification of single materials to‌ an adaptive design of the entire battery system. This review assesses the research conducted on NVPF cathodes over the past three years from several perspectives, focusing on the feasible application of materials over a wide temperature range at high voltages, summarizing the challenges and required development strategies in future research.

Keywords

Sodium-ion battery, sodium vanadium fluorophosphate, cathode materials synthesis, materials modification, battery optimization

INTRODUCTION

The negative environmental impact associated with the use of fossil fuels has driven the development of alternative renewable energy technologies, such as solar and wind power[1], which has prompted significant research activity with respect to efficient electrochemical energy storage technologies for practical application[2,3]. Lithium-ion batteries (LIBs) have emerged as viable options, exhibiting low self-discharge rates and high energy density[4-6]. However, the widespread application of LIBs has been limited by the high cost and uneven distribution of lithium resources[7-9]. In this context, sodium-ion batteries (SIBs) have been the subject of appreciable research as they are not subject to resource constraints, and can be operated safely at a low cost[10-12]. As a homologous element, sodium possesses similar physicochemical properties to lithium, and some of the technologies and strategies established for LIBs can be applied to SIBs. Moreover, inexpensive aluminum foil can be used as the anode current collector in SIBs, reducing associated costs[13,14]. However, SIBs exhibit low Na+ transport rates, low energy density, and a limited reversible capacity[15]. Therefore, it is necessary to improve SIB cycle stability and energy density in order to enable effective application, which requires significant scientific research and engineering technology advancements.

The electrochemical performance of SIBs is dependent on the inherent chemical properties of the electrode materials[16-18]. Moreover, the cathode materials largely determine the device mass and output voltage[19-23]. Currently, the SIB cathode materials can be divided into three categories: transition metal oxides[24-28], Prussian blue analogs[29-40], and polyanionic compounds (PACs)[41-45]. Transition metal oxides exhibit a high specific capacity and facile synthesis, but the associated low average voltage, low energy density, poor cycle stability, and humidity sensitivity must be addressed[46]. Prussian blue analogs have an open framework, abundant redox active sites, good structural stability, and large ion channels, but the coulombic efficiency and cycle stability are affected by lattice defects and crystal water[47,48]. PACs are characterized by superior structural and thermal stability with enhanced ion diffusion dynamics that contribute to cycle stability and safety[49,50]. In particular, Na3V2(PO4)2F3 (NVPF) as a sodium (Na) super ionic conductor (NASICON), has been reported to deliver a high output voltage (3.95 V) and energy density (up to 500 Wh kg-1) as a result of the highly ionic F-V bond, which enables an electrochemical output comparable to lithium iron phosphate-based LIBs[51,52]. However, there are a number of challenges in terms of NVPF practical application, including the low intrinsic conductivity, high structural stress during electrochemical processes, and the high cost of vanadium[53]. In order to improve electrochemical performance and reduce costs, it is necessary to adopt effective strategies for enhancing conductivity and Na+ diffusion kinetics, and establish an effective commercially viable vanadium alternative.

Previous reviews predominantly focused on the intrinsic properties and singular modifications of NVPF[53-55], primarily confined to laboratory-scale optimizations, while insufficient attention was paid to the extreme operational conditions and cost-effectiveness. This review constructs a multi-scale collaborative research framework “atom-interface-system”, breaking through the conventional singular emphasis on intrinsic material performance in the existing reviews. By incorporating high-entropy design, dynamic failure mechanisms, extreme-condition adaptability, and scalable manufacturing into a systematic analysis, we propose a reference technical roadmap to advance the industrial application of NVPF. The existing challenges in NVPF research are noted and future developments are outlined with the goal of realizing the practical application of high-performance NVPF systems.

CRYSTAL STRUCTURE AND PROPERTIES OF NVPF

Sodium vanadium fluorophosphate exhibits a typical NASICON structure, consisting of a three-dimensional (3D) framework formed by [V2O8F3] double octahedra and [PO4] tetrahedra linked through oxygen atoms. In this structure, the double octahedral units are connected by fluorine atoms, whereas the tetrahedral units are connected by oxygen atoms. The presence of [PO4] facilitates rapid migration of Na+ within the framework and stabilizes the structure during metal redox processes.

Structural analysis of NVPF reveals minor orthorhombic lattice distortions localized on the (h00) and (0k0) crystallographic planes, with Rietveld refinement confirming the space group as Amam[56]. The Na+ ions occupy three distinct crystallographic sites (labeled Na1, Na2, and Na3) with refined occupancy ratios of 1:4/3:2/3, respectively [Figure 1A]. In a subsequent study, the desodiation process of Na3V2(PO4)2F3(Na3VPF) was elucidated by in situ X-ray diffraction (XRD) analysis[57]. As shown in Figure 1B, (Na3VPF) (space group Amam) initially forms distinct intermediate phases during desodiation, including Na2.4VPF, Na2.2VPF (I4/mmm), Na2VPF (I4/mmm), and Na1.8-1.3VPF (I4/mmm), before ultimately transforming into NaVPF (Cmc21). In this process, the space group transitions sequentially from Amam to I4/mmm and finally to Cmc21. This structural evolution enables a two-electron transfer reaction, delivering a theoretical specific capacity of approximately 128 mAh g-1. The structural evolution of NVPF and the sodium insertion/extraction processes are expressed in[58]:

Recent research progress in the application of Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>2</sub>F<sub>3</sub> cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization

Figure 1. (A) Crystal structure of NVPF with the Amam space group;(B) Corresponding crystal structural evolution of NVPF during Na+ extraction. Reproduced with permission[57]. Copyright 2015, American Chemical Society; (C) The key benefits and intrinsic challenges of NVPF cathodes in SIBs. NVPF: Na3V2(PO4)2F3; SIB: sodium-ion battery.

$$ \mathrm{Na}_{3} \mathrm{~V}_{2}^{3+}\left(\mathrm{PO}_{4}\right)_{2} \mathrm{~F}_{3} \leftrightarrow \mathrm{Na}_{2} \mathrm{~V}^{3+} \mathrm{V}^{4+}\left(\mathrm{PO}_{4}\right)_{2} \mathrm{~F}_{3}+\mathrm{Na}^{+}+\mathrm{e}^{-} \\ $$

$$ \mathrm{Na}_{2} \mathrm{~V}^{3+} \mathrm{V}^{4+}\left(\mathrm{PO}_{4}\right)_{2} \mathrm{~F}_{3} \leftrightarrow \mathrm{NaV}_{2}^{4+}\left(\mathrm{PO}_{4}\right)_{2} \mathrm{~F}_{3}+\mathrm{Na}^{+}+\mathrm{e}^{-} \\ $$

It should be noted that NVPF suffers from intrinsically poor electronic conductivity (~ 10-12 S cm-1) due to the insulating [PO4] tetrahedra that spatially separate V atoms and impede electron transport. Moreover, repeated Na+ insertion/extraction induces significant structural stress and anisotropic volume fluctuations, which serve to degrade mechanical stability and lead to a severe loss of capacity during cycling. The advantages and intrinsic challenges associated with NVPF are illustrated schematically in Figure 1C.

In the synthesis of NVPF, factors such as the carbon source selection and reaction conditions (temperature and pH) may result in fluorine loss‌. Fluorine deficiency disrupts the corner-sharing connectivity of [PO4] tetrahedra in the NASICON framework, destabilizing the 3D ion channels. This structural distortion triggers localized stress accumulation and significantly lowers the Na+ migration rate[59]. At a low fluorine content, irreversible phase transitions (such as the transition from high-pressure to low-pressure regions) occur during the charge/discharge process, accelerating structural collapse[60]. Moreover, the corrosion acid (HF) induced by trace H2O at the NVPF/Al collector interface further causes fluorine loss, degradation of crystal structure, increase in polarization voltage, and shedding of some active materials, leading to a rapid capacity decline[61]. Fluorine loss can be effectively mitigated by optimizing the carbon source, process temperature, and anion regulation.

NVPF SYNTHESIS

A number of NVPF synthesis procedures have been reported, including solid-state[62-66], sol-gel[67-70], spray drying[71-73], hydrothermal/solvothermal[74-76], and electrostatic spinning[77-79] methods. In addition, carbon coating, calcination, microwave treatment, and auxiliary reagents have been considered.

Solid-state method

The solid-state procedure is easy to operate with a low associated cost, typically involving mechanical grinding and sintering[80]. An increase in calcination temperature and the use of nanoscale raw materials can improve product purity and reduce overall reaction time.

Minart et al.[81] reported dense and crystalline carbon-coated Na3V2(PO4)2F3 (NVPF@C) nanoparticles resulting from solid-state synthesis, with a tap density of 1.4 g cm-3 (increased by 40%) following mechanical grinding, which represents the highest published value. The mechanical processing reduced the size of the particles and crystals while retaining the NVPF structure and surface chemical properties.

In the case of carbon thermal reduction (CTR), the carbon source is a critical consideration, serving as a reducing agent and coating layer to improve NVPF electronic conductivity. A number of organic carbon (glucose[66] and bitumen[64]) and inorganic carbon (acetylene black[82] and super-phosphorus[83]) sources have been considered. Different carbon sources were mixed with V2O5 and NH4H2PO4 by ball-milling, and sintered at 800 °C to prepare carbon-coated NVPF materials (C-NVPF)[84]. Samples prepared using glucose as the sole carbon source delivered the best electrochemical performance in SIBs. Tannic acid has been employed as a chelating agent and carbon source to prepare a carbon-coated composite [NVPF composite coated by tannic acid-derived carbon (NVPF@C-T)], using a combination of solid-state and sol-gel methods[85]. By introducing tannic acid in two stages, a comprehensive and consistent carbon coating layer can be successfully created which significantly enhances the electronic conductivity of NVPF, facilitates the diffusion of Na+, and safeguards NVPF from corrosion caused by the electrolyte.

Citric acid was used as both a reducing agent and a carbon source in the synthesis of C-NVPF using surface ball-milling [Figure 2A][86]. The carbon coating improved conductivity and inhibited the aggregation of nanoparticles. The low voltage platform and decreased energy density due to the formation of NVP can be addressed by employing ball-milling to substitute Al for V in preparing NVAPF[87]. The partial substitution of Al for V alters the local electronic states, enhancing the stability of F, where kinetic studies have confirmed an increase in Na+ migration rate.

Recent research progress in the application of Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>2</sub>F<sub>3</sub> cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization

Figure 2. (A) Schematic representation of carbon coated NVPF synthesis by facial ball-milling. Reproduced with permission[86]. Copyright 2024, Royal Society of Chemistry; (B) Schematic representation of rectangular NVPF synthesis using a hydrothermal method. Reproduced with permission[90]. Copyright 2023, Wiley -VCH; (C) SEM image of the materials synthesized by the sol-gel technique. Reproduced with permission[97]. Copyright 2024, Royal Society of Chemistry; (D) Schematic representation of NVPF composite synthesis by spray drying. Reproduced with permission[102]. Copyright 2024, Wiley; (E) SEM image of the materials synthesized by electrostatic spinning. Reproduced with permission[105]. Copyright 2023, American Chemical Society; (F) SEM image of the materials synthesized using one-step chemical vapor replacement. Reproduced with permission[107]. Copyright 2024, Wiley -VCH. NVPF: Na3V2(PO4)2F3; SEM: scanning electron microscope.

Although the NVPF product generated by solid-state synthesis has a high phase purity, the electrochemical performance is affected by poor compositional homogeneity and a wide particle size distribution. It is necessary to further optimize the solid-state synthesis procedure in order to obtain NVPF with a uniform composition.

Hydrothermal/solvothermal method

Applying a hydrothermal/solvothermal synthesis enables materials synthesis to generate a product that effectively alleviates structural stress during electrochemical operation. For instance, a microwave-assisted hydrothermal procedure has been used to uniformly distribute NVPF on reduced graphene oxide (rGO), where NVPF nanocuboids coated with rGO (NVPF@rGO) were obtained after calcination[88]. Al-Marri[89] reported the synthesis of NVPF/rGO with a 3D conductive structure using a simple hydrothermal/solvent thermal reduction process. With the help of rGO, NVPF/rGO exhibited improved surface kinetics and diminished polarization, ultimately resulting in enhanced rate performance and extended cycle stability. The application of phosphomolybdic acid (PMA) as a crystal plane inducer served to increase adsorption and surface energy, accelerating the growth of NVPF crystals, and resulting in rectangular NVPF with continuous ionic diffusion paths. A nitrogen-doped carbon-coated rectangular NVPF cathode material (c-NVPF@NC) was synthesized by a PMA-assisted hydrothermal method [Figure 2B][90].

The grain size and vacancy defects can be effectively adjusted by varying the annealing temperature. For example, a NVPF electrode material obtained using hydrothermal method and subsequent thermal annealing at 350 °C exhibited a specific capacitance of 167.73 F g-1 and a capacitance retention of 85%[91]. Lin et al.[92] studied the crystallization of NaH2PO4-VOSO4-NaF-H2O under hydrothermal conditions, where Na3V2O2x(PO4)2F3-2x was produced at 170 °C after 2 h with vanadium present in mixed valence states (V3+ and V4+). Varying the synthesis time, vanadium source, or the addition of reducing agents resulted in the formation of different NVPF pure phases. The unique pore structure of the NVPF cathode offers more active sites and promotes the diffusion of Na+. For example, tNVPF@C with a dense microsphere morphology was prepared by sodium dodecyl benzene sulfonate (SDBS)-assisted solvothermal treatment, exhibiting a stable structure, high conductivity and large specific surface area[93]. By carbonizing polyacrylamide (PAM) to create a coating, the electronic conductivity of tNVPF was significantly enhanced, resulting in tNVPF@C with exceptional Na+ storage capabilities. In addition, a 3D porous carbon was employed as a nucleation substrate to synthesize two-dimensional (2D) NVPF nanosheets (NS) [Na3V2(PO4)2F3 nanosheets on 3D porous carbon (NVPF@3Dc)] in a one-step hydrothermal method[58]. A rapid ion transfer within the a-b plane of the NVPF@3Dc crystal, combined with a minimal volume change during Na insertion and extraction, contributed to enhanced electrode rate performance.

Hydrothermal/solvothermal methods can enhance reaction rates and generate specific structures, offering significant application potential. However, the synthesis results in variable product morphology and performance. Further research is required to tune the procedure in order to control material uniformity and stability, particularly in large-scale synthesis of NVPF.

Sol-gel method

The sol-gel method represents a simple process, requiring a relatively low reaction temperature, and is widely applied in materials synthesis[94,95]. Sun et al.[96] have reported a dual-functional N-doped carbon and β’’-Al2O3 coated NVPF (NC/β’’@NVPF) using the sol-gel procedure with subsequent calcination. The NC/β’’@NVPF cathode exhibited a high capacity (126.7 mAh g-1) at 0.2 C, maintaining 95.5% of cpacity after 500 cycles (at 10 C). This response was attributed to compensation for the loss of Na+ by Na+ in β’’-Al2O3 during the formation of the cathode/electrolyte interface film.

Citric acid and oxalic acid are typically used as cross-linking agents and carbon sources in sol-gel synthesis. Yang et al.[97] prepared a carbon-coated Na3V2(PO4)2F3·3Na3V2(PO4)3 composite (NVPF·3NVP/C) composite [Figure 2C] where citric acid was employed to fabricate a 3D conductive carbon layer. The synergistic effect of the composites facilitated rapid ionic migration and enhanced the kinetic reactions. Notably, NVP, serving as the primary phase, reinforced structural stability and substantially boosted the capacitive contribution. In a separate study, citric acid was utilized as a chelating agent to synthesize NVPF by sol-gel[98] where an excess or insufficient chelating agent resulted in F atom deficiency with a loss of crystallinity. It was found that the specific capacity, rate capability, and cycling performance were significantly influenced by the crystallinity of NVPF.

Recent work has considered synergistic strategies involving sol-gel treatment directed at surface modification. For instance, Mahato et al.[99] have provided the first report of rota-tumbler assisted sol-gel to synthesize a series of carbon-coated Fe and Cr co-doped NVPF (NVFCPF). An enlargement in the size of the saddle point, along with the planes of monocapped trigonal prisms and the hexagonal cavities, aided in reducing the ion diffusion activation energy, thereby enhancing the diffusivity of the doped material. In addition, a carboxymethyl cellulose (CMC)-assisted sol-gel strategy was utilized to prepare cross-linked carbon-coated porous NVPF composites (NVPF@C@CMC)[100]. The CMC provided additional Na+ to compensate for sodium loss during the formation of the solid electrolyte interface. Structural stability was enhanced by the cross-linked carbon, and the porous structure of the coating facilitated Na+ transfer.

Although the sol-gel method can be used to prepare porous NVPF, additional reaction conditions should be investigated to further improve material pore structure and morphology, such as pH, precursor composition, and temperature.

Spray drying

Spray drying method is commonly used to prepare powder particles or films, offering notable commercial benefits, such as low cost, short processing time, and ease of production. Different morphologies and structures can be achieved by adjusting the preparation conditions and employing auxiliary reagents. For example, a Br-doped spherical NVPF/C was prepared using polytetrafluoroethylene (PTFE) and cetyltrimethylammonium bromide (CTAB) as F supplements and regulators, respectively, in a one-step spray drying procedure[101]. During the synthesis process, the hard PTFE template supplied additional F, and the resultant porous structure facilitated electrolyte permeation. The soft CTAB template provided Br, and the self-assembly process resulted in a strong spherical structure while refining the particle size. In another study, Zhou et al.[102] synthesized copper substitution NVPF microspheres with an eggshell structure using polyvinyl pyrrolidone (PVP)-assisted spray drying [Figure 2D]. The high electronic conductivity of the cathode material and rapid Na+ diffusion were ascribed to the unique eggshell structure and crystal lattice.

Carbon-coated NVPF/carbon nanotube (CNT) nanospheres were synthesized by spray drying followed by calcination[103]. The 3D CNT network and carbon coated layer enhanced electronic conductivity, increasing the contact area between the cathode and electrolyte, and improving overall the structural stability. Applying sugarcane bagasse as raw material, carbon-coated NVPF nanospheres (NVPF/C) with high capacity were prepared by spray drying combined with high-temperature calcination[104].

Although the NVPF particles prepared by spray drying exhibited a high degree of uniformity, the structures reported to date are relatively simple, and the ability to control structure by adjusting preparation conditions warrants further investigation.

Electrostatic spinning

Electrostatic spinning involves a relatively simple operation that produces materials with a uniform morphology, and is particularly suitable for synthesizing materials with a one-dimensional linear/tubular structure. These structures can facilitate an oriented transmission of electrons, enhancing reaction kinetics via a conductive network. Electrodes prepared by electrostatic spinning possess a self-supporting structure that serves to increase the specific energy by reducing the use of binders and conductive carbon.

Sodium storage may be improved by adjusting spinning parameters, combined with high-temperature calcination, hydrothermal synthesis, chemical deposition, and other techniques. The porous structure and tubular characteristics can provide channels for Na+ transfer, enhancing conductivity. Liang et al.[105] utilized electrostatic spinning to prepare self-standing NVPF cathodes [Figure 2E]. The NVPF nanoparticles were evenly distributed on the surface and inside the carbon nanofibers (CNFs), which effectively limited particle volume expansion during charging and discharging. Moreover, the CNF network facilitated electron and ion transport. The NVPF self-standing cathode exhibited a high-voltage platform of ~ 4.07 V, and the assembled full cell demonstrated an energy density as high as 185.6 Wh kg-1.

Electrostatic spinning requires stringent experimental conditions as the process is sensitive to temperature and humidity, and involves extended preparation times which have limited large-scale application. In addition, the inorganic fibers produced are susceptible to damage where the fiber length may be significantly decreased during grinding.

Other methods

A number of researchers have considered alternative routes to NVPF materials synthesis. In order to improve the low tap density and poor cycling performance, Song et al.[106] synthesized NVPF materials (HTS-NVPF) with a uniform conductive network and high tap density using a high-temperature shock (HTS) strategy. The ultra-fast high-temperature synthesis (heating rate = 1,100 °C s-1 and heating time = 10 s) circumvented F loss and the formation of impurities, while enhancing cycling stability. The conductive network associated with the large synthesized particles was uniform, contributing to increased conductivity. The HTS-NVPF system exhibited a wide range of operating temperatures (-45 °C to 55 °C).

In order to address Na+ diffusion rate, Wang et al.[107] prepared Na3-yVPO2-xBrxF by doping Br- and pre-loading Na+ vacancies using a one-step chemical vapor replacement (CVR) method [Figure 2F]. The electrostatically shielded channels in Na3-yVPO2-xBrxF reduced coulombic hindrance by Na+ during the insertion/extraction process, increasing Na+ diffusion rate by a factor of 5.

A flexible cathode with a high-rate capability and stability can satisfy the requirements of energy storage devices with respect to portability, high power and energy density. A flexible cathode material without binder (NVPF@C/CC) was prepared using a colloidal crystal template method, which anchored nanocarbon onto carbon cloth that was used to coat the NVPF, providing a 3D ordered microporous structure[108]. The latter facilitated ion diffusion, while the combination of the carbon layer and carbon cloth provided a continuous electronic transmission path, which contributed to a high-rate performance and conductivity.

The electrochemical properties of the NVPF materials synthesized using a range of methods are presented in Table 1.

Table 1

Electrochemical performance of NVPF cathodes in SIBs prepared using different procedures

CompositePreparation methodElectrolyteElectrochemical performances
(current density, discharge capacity, cycles, and capacity retention)
Ref.
NVPF@Csolid-state methodEC and DMC (1:1) 20 C, 123 mAh g-1, 100, 87.8% (5 C)[81]
NVPFcarbon thermal reduction method1 M NaClO4 in PC & 5 wt.% FEC0.1 C, 122 mAh g-1, 1,000, 83.2% (10 C)[84]
NVPF@C-Tsolid-state method1 M NaClO4 in EC: DMC: EMC = 1:1:1 with 2 vol.% FEC0.2 C, 124.5 mAh g-1, 300, 94.8% (10 C)[85]
C-NVPFsurface ball milling method1 M NaClO4 in EC/DEC (1: 1, by vol)0.1 C, 110.6 mAh g-1, 2,000, 54% (5 C)[86]
NVAPFball milling method1 M NaClO4 in PC0.5 C, 117.4 mAh g-1, 100, 92.4% (0.5 C)[87]
NVPF@rGOhydrothermal method1 M sodium NaPF6 in diglyme1 C, 123 mAh g-1, 7,000, 85.6% (10 C)[88]
NVPF/rGOHydrothermal /solvent thermal1 M NaPF6 in EC: DEC: DMC (1: 1:1, by vol)0.1 C, 127 mAh g−1, 100, 99% (25 C)[89]
c-NVPF@NChydrothermal method1 M NaClO4 in EC: DMC: EMC (1:1:1) with 5 vol.% FEC0.2 C, 121 mAh g-1, 1,000, 78% (10 C)[90]
Na3V2(PO4)2F3hydrothermal method 2 M NaOH1 A g-1, 168 F g-1, 800, 85% (1 A g-1)[91]
solvothermal method1 M NaClO4 in EC: DMC: EMC (1:1:1) with 2 vol.% FEC0.2 C, 120.7 mAh g-1, 300, 80% (5 C)[92]
NVPF@3Dchydrothermal method1 M NaClO4 in EC: PC (1:1) with 2 vol.% FEC0.2 C, 131.5 mAh g-1, 2,500, 64% (5 C)[58]
NC/β’’@NVPFsol-gel method1 M NaClO4 in EC: DMC: EMC (1:1:1) with 2 vol.% FEC0.2 C, 126.7 mAh g-1, 500, 95.5% (10 C)[96]
NVPF·3NVP/C sol-gel method1 M NaClO4 in EC: DEC (1:1, by vol) 0.1 C, 120.7 mAh g-1, 200, 64.1% (10 C)[97]
NVPFsol-gel method1 M NaClO4 in EC: DEC (1:1, by vol) with 5 vol.% FEC0.2 C, 112.9 mAh g-1, 1,000, 84.6% (5 C)[98]
NVFCPFsol-gel method1 M NaClO4 in EC: PC (1:1) with 2 vol.% FEC0.1 C, 120 mAh g-1, 500, 88.6% (1 C)[99]
NVPF@C@CMCsol-gel method1 M NaClO4 in EC: DMC: EMC (1:1:1) with 5 vol.% FEC0.2 C, 125.8 mAh g-1, 200, 77% (1 C)[100]
NVPF/Cspray drying method1 M NaClO4 in EC: PC (1:1) with 5 vol. % FEC1 C, 116.1 mAh g-1, 1,000, 98.3% (10 C)[101]
NVPF@NC-Cuspray drying method1 M NaClO4 in EC: PC (1:1) with 5 vol. % FEC0.1 C, 117.4 mAh g-1, 5,000, 91.3% (20 C)[102]
NVPF/CNTsspray drying method1 M NaClO4 in EC: PC (1:1) with 5 vol. % FEC0.1 A g-1, 127.7 mAh g-1, 20000, 92.1% (10 A g-1)[103]
Na3V2(PO4)2F3/Cspray drying method1 M LiPF6 in EC: DMC (1: 1, by vol)0.1 C, 125 mAh g-1, 80, 78.23% (0.1 C)[104]
NVPFelectrostatic spinning method1 M NaClO4 in K1: EC: PC (1:3:6) with 2 vol.% FEC0.2 C, 101.8 mAh g-1, 400, 98.3% (0.2 C)[105]
HTS-NVPFhigh-temperature shock strategy1 M NaClO4 in EC: PC with 5 vol.% FEC0.1 C, 116.8 mAh g-1, 200, 94.2% (1 C)[106]
Na3-yVPO2-xBrxFchemical vapor replacement method1 M NaClO4 in EC: PC (1:1) with 5 vol.% FEC0.2 C, 124.6 mAh g-1, 800, 94.4% (10 C)[107]
NVPF@C/CCcolloidal crystal template method1 M NaClO4 in EC: DMC (1: 1, by vol)0.1 C, 118.2 mAh g-1, 2,000, 83.3% (20 C)[108]

NVPF MODIFICATION STRATEGIES

It is possible to construct channels for ion transport by regulating the NVPF morphology, which effectively enhances electrochemical performance. Such regulation also serves to reduce the impact of stress on the material volume, and minimize secondary reactions when the electrode contacts the electrolyte. At the nanoscale, the number of active sites on the electrode is increased and the distance for ion diffusion is reduced, which promotes Na+ diffusion, increasing overall electrochemical activity. The electrochemical properties of materials can be precisely tailored through appropriate modification strategies. Currently, the main modification strategies include carbon coating[109,110], the use of carbon material composites[111-114], and elemental doping[115-124].

Carbon coated NVPF

One of the key approaches to optimizing electrode structure is to construct an electron transfer system with high electrical conductivity. It is possible to effectively enhance ion and electron conductivity, and increase electrode stability using surface coating techniques[125]. Carbon materials are commonly used in coating, offering high conductivity and porous characteristics that provide effective transport channels, increasing electrochemical reaction rates while effectively avoiding direct contact with the electrolyte, which mitigates volume changes during Na+ extraction/insertion[126]. Sun et al.[127] prepared carbon-coated NVPF composites using the sol-gel method, employing PTFE as the carbon source and a fluorine supplement. The presence of the NVP impurity in these carbon-coated NVPF composites was markedly diminished or entirely eradicated.

Carbon-based materials can control NVPF morphology and structure during synthesis and sintering, providing a scaffold that reduces the stress impact caused by electrical effects. Hu et al.[128] have prepared microsphere composite materials (HM-NVPF@CN) assembled from carbon-coated NVPF NS [Figure 3A]. Subsequently, a cathode-electrolyte interface (CEI) layer rich in B and F was constructed on the surface. The unique 2D NS structure of HM-NVPF@CN can shorten the diffusion paths of Na+ and e-, increasing the reaction contact area while inhibiting parasitic reactions. The thin carbon coating is beneficial for improving electronic conductivity and mitigating structural stress. The optimized electrode exhibited superior cycling stability over a wide temperature range (-25 °C to 60 °C). In addition, NVPF with a porous microsphere structure was synthesized using the deep eutectic solvent method[129]. The composite carbon coated Na3V2(PO4)2F3 microspheres with fluorocarbon layer NVPF/C-F was prepared by forming an amorphous fluoroarbon layer via PTFE surface modification. The study demonstrated that the amorphous fluoroarbon layer promoted Na+ diffusion and charge transfer.

Recent research progress in the application of Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>2</sub>F<sub>3</sub> cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization

Figure 3. (A) Finite element simulation models and design strategies for NVPF microspheres assembled using carbon coated NVPF nanosheets. Reproduced with permission[128]. Copyright 2024, Elsevier B.V.; (B) Schematic representation of materials preparation; (C) TEM images of a dual-function carbon network composed of nitrogen-doped carbon layers and carbon bridges. Reproduced with permission[133]. Copyright 2024, American Chemical Society. NVPF: Na3V2(PO4)2F3; TEM: transmission electron microscope.

Layered carbon materials provide a bridge between electrons and active substances, and effectively promote the migration of Na+ between the electrolyte and electrodes. Carbon-coated NVPF with a microcube structure (NVPF/C) was prepared using an open system synthesis method, involving in situ growth on graphene oxide (GO) microcubes[130]. The NVPF/C has an average particle size of ~ 2 μm and a carbon layer thickness of ~ 5 nm. Due to the structure-induced electrochemical reversibility and improved reaction kinetics, it demonstrated the superior electrochemical properties. The carbon coatings prepared in situ from inorganic or organic carbon sources exhibit different structural characteristics. For example, carbon-coated NVPF composites were synthesized using branched starch as carbon source and structural template[131]. Two-dimensional nanostructures were formed where the branched starch induced NVPF nucleation along its main chain. This structural characteristic can shorten the Na+ diffusion path and enhance electron transfer.

Carbon coating doped with heteroatoms can significantly improve electronic conductivity and structural stability where the dopants introduce numerous active sites and structural defects, which enhance ion transport efficiency between the electrode and electrolyte. N-doping provides abundant active sites for electrochemical reactions and enhances surface hydrophilicity. An N-doped carbon-coated NVPF composite was prepared using PVP as nitrogen source using the sol-gel method[132]. The N doping introduced exogenous defects in the carbon layer and additional active sites, increasing conductivity and promoting Na+ diffusion.

It is known that dopamine-based coatings can spontaneously accumulate on the surface of the materials. Sun et al.[133] fabricated a bifunctional carbon network by pyrolysis of dopamine (PDA) and PTFE, encapsulating NVPF particles and carbon bridges with an N-doped carbon layer [Figure 3B and C]. The PTFE served to supplement F loss during high-temperature calcination, inhibiting the formation of NVP impurities. The dopamine provided N for the carbon coating layer, creating defects and additional Na storage sites. The bifunctional N-doped carbon network can optimize electron/ion transfer at the interface of the electrode material, preventing NVPF corrosion by the electrolyte.

When compared with single atom doping, double atom doping can introduce more active sites, effectively promoting electron and ion transport[134]. This is particularly the case for doping with N, S, and B, which serves to improve Na+ storage capacity and electronic conductivity in diverse modification processes[135]. In a recent study, a N, B-doped carbon-coated NVPF composite (NVPF@NBC) was prepared by a mechanical ball-milling assisted sol-gel procedure, employing polyacrylonitrile (PAN) and boric acid as nitrogen and boron sources, respectively[136]. The co-doping of N and B atoms increased the number of defects in the carbon, enhancing Na+ storage and transport. The N, B co-doped carbon layer can reduce particle agglomeration and maintain structural stability. Yu et al.[137] synthesized a N/B co-doped carbon coated NVPF (NVPF-CNB) using the sol-gel method with 1-vinyl-3-methyl imidazole tetrafluoroborate as the nitrogen and boron source, and a carbon-coated NVPF composite using ionic liquid-acrylic copolymer as a modification reagent[138]. A N/S co-doped carbon-coated NVPF (NVPF@NSC) was also prepared by sol-gel combined with mechanical ball-milling which employed thiourea as the nitrogen and sulfur source[139]. The study revealed that the doping of N and S had no impact on the crystal structure of NVPF, but demonstrated excellent reversibility and minimal polarization.

Despite the widespread use of carbon-coated NVPF composite materials, graphitization in carbon layers remains a challenge that impairs electronic conductivity. Moreover, it is difficult to form uniform carbon layers using current preparation procedures, which has a significant impact on Na+ migration between particles and electrolytes. It should also be noted that carbon layers formed from different carbon sources have varying electrochemical properties. Preparation conditions and the use of regulatory factors warrant further investigation.

Carbon composite NVPF

When NVPF particles experience fracture or loss of potential contact, the resistance between them intensifies, resulting in decreased capacity. It is essential to construct a stable NVPF structure and maintain a consistent and reliable electrochemical output. Three-dimensional conductive carbon materials, notably CNTs, are recognized as the most suitable candidates for this purpose. Research has conclusively demonstrated the efficacy of CNTs in boosting ionic transport[140]. By embedding CNTs within NVPF particles and constructing an interconnected network, the material can better accommodate Na+ volumetric fluctuations during repeated charging and discharging, maintaining the electrical conductivity between self-assembled particles. For example, Ma et al.[141] synthesized a CNT-hybrid dual-phase anionic phosphate material (denoted as NFC) via a one-pot sintering method. The material consisted of NVPF with interwoven CNTs and a minor amount of NVP. This cathode material demonstrated a wide operating temperature range (-20 °C to 50 °C). A NVPF@CNT composite was prepared by anchoring NVPF nanoparticles on CNTs with a low number of walls [Figure 4A-C][142]. Ion transmission via the CNTs promoted the growth of primary NVPF crystal particles. The high crystallinity and anchoring effect enhanced structural stability, while the 3D network composed of CNTs and amorphous carbon layers increased conductivity. This strategy can improve material crystallinity and increase grain size. CNTs were utilized as an effective medium to adjust the electronic conductivity of NVPF in a recent study[143].

Recent research progress in the application of Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>2</sub>F<sub>3</sub> cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization

Figure 4. (A) Schematic representation of ion transport by CNTs to promote NVPF grain growth; (B) Synthesis process of NVPF@CNT composite materials; (C) Cycle performance of different materials. Reproduced with permission[142]. Copyright 2023, Elsevier; (D) SEM and (E) HRTEM images of NVF-GEL/rGO. Reproduced with permission[149]. Copyright 2022, Elsevier. (F) Schematic representation of the synthesis process. The illustration is TEM image of NVPF@rGO/CNTs. Reproduced with permission[150]. Copyright 2023, Wiley. CNT: Carbon nanotube; NVPF: Na3V2(PO4)2F3; NVPF@CNT: Na3V2(PO4)2F3 coated with carbon nanotubes; SEM: scanning electron microscope; HRTEM: high resolution transmission electron microscope; NVF-GEL/rGO: NVPF modified reduced graphene oxide using GEL method.

Research has shown that the formation of a 3D cross-linked carbon network can enhance charge transfer kinetics and shorten the ionic transfer pathway. For instance, a composite with a cross-linked dual-carbon framework structure [Na3V2(PO4)2F3 caged in 3D cross-linked carbon networks composed of carbon layer and intertwined CNTs (NVPF@C@CNTs)] was constructed by cross-linking NVPF with amorphous carbon and CNTs[144], and NVPF@C/CNTs composites with a carbon-coated 3D conductive network were prepared by solvothermal method [145].

Introducing graphene is an effective strategy to improve electrode performance. GO can serve as an efficient carrier for immobilizing nanoparticles due to its unique 2D structure[146,147]. A high-temperature heat treatment is used to prevent particle agglomeration and lower the GO content with a stable anchoring of NVPF, and an extended contact area between the electrode and electrolyte[148]. For instance, NVPF-gel/rGO composites were synthesized using a hydrothermal procedure with addition of gelatin and rGO[149]. The morphology of the material exhibited a high degree of uniformity where the addition of gelatin reduced impurities, and the presence of graphene enhanced Na+ diffusion [Figure 4D and E]. Shi et al.[150] prepared a composite material (NVPF@rGO/CNT) by self-assembly, generating a 3D cross-linked conductive network containing 2D rGO and 1D CNT [Figure 4F], which makes it feasible to operate in a wide temperature range (-40 °C to 50 °C).

Combining conductive carbon with NVPF is an effective strategy to enhance electronic conductivity. This is particularly the case when the carbon materials form 3D or dual-carbon layer wrapping structures.

Element doping of NVPF

NVPF materials typically exhibit lower intrinsic electronic conductivity and limited rate performance due to the insulating properties of the [PO4] tetrahedron. Element doping enables a flexible adjustment of material properties, such as electronic conductivity, structural defects, crystallinity, and redox potential. The doping elements can be classified into three categories based on differences in the substitutional sites: sodium ion site substitution; vanadium ion site substitution; anion site substitution. Recent research has mainly focused on the first two types of substitution.

Doping elements with a larger ionic radius at the Na site in NVPF materials can lead to an increase in lattice spacing that, in turn, widens the channels for ionic diffusion[151,152]. This ensures that the NASICON framework is not damaged or affected by internal stress and structural deformation during charge and discharge cycles. Moreover, the introduction of these elements results in a shortening of the vanadium-fluorine bond (V-F) and vanadium-oxygen bond (V-O) bond lengths in the lattice, enhancing NVPF structural stability. Wu et al.[153] have prepared composite materials with a carbon coating by the sol-gel method, incorporating K+ to replace a Na+ component in NVPF crystals [carbon coated Na3V2(PO4)2F3 doped with K+ (N1-xKxVPF/C)]. The doped K+, which was located at the Na1 site, did not participate in reaction during charge-discharge, ensuring high Na+ migration. The doping of K+ reduced the bandgap energy and increased electronic conductivity.

The introduction of high-entropy elements (Mg, Cr, Al, Nb) can improve the stability of the Na2 site in NVPF, and suppress the intermediate reaction of the NVPF phase at low potentials. High-entropy carbon-coated NVPF composites (NVPF-HE@C) [Figure 5A], synthesized by the sol-gel method, exhibited a significant improvement in the specific capacity, capacity retention, and Na+ diffusion coefficient[154]. Doping with Fe serves to refine the particles, inhibiting particle agglomeration, and promoting Na+ diffusion. Yang et al.[155] prepared NVPF composites [Fe-doped Na3V2(PO4)2F3 coated by N-doped CNTs (Fe-NVPF@N-CNTs)] composed of Fe-doped NVPF and N-doped CNTs, which exhibited enhanced conductivity due to a structure of 3D carbon framework.

Recent research progress in the application of Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>2</sub>F<sub>3</sub> cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization

Figure 5. Schematic representation of ion-doped NVPF crystal structures: (A) High-entropy element. Reproduced with permission[154]. Copyright 2024, Elsevier; (B) Cr3+. Reproduced with permission[166]. Copyright 2024, Elsevier; (C) Mn2+. Reproduced with permission[169]. Copyright 2023, MDPI; (D) Ti4+. Reproduced with permission[170]. Copyright 2023, Chemistry Europe; Rate performance of ion-doped NVPF materials: (E) Ca2+. Reproduced with permission[162]. Copyright 2024, American Chemical Society; (F) Al3+. Reproduced with permission[15]. Copyright 2022, Elsevier; (G) Cr3+. Reproduced with permission[166]. Copyright 2024, Elsevier. NVPF: Na3V2(PO4)2F3.

When compared to doping at the Na site, the method of replacing V with inexpensive elements represents a more common approach that can reduce the toxicity due to V and lower overall costs. Inactive elements (such as Mg2+[156,157], Al3+[158,159], and Y3+[117]) that do not participate in electrochemical reactions, can promote the formation of defects in the NVPF crystals, providing additional charge transfer pathways.

The holes that are generated by the substitution of V by Mg enhance conductivity. The Mg2+ dopant does not participate in redox reactions, but can suppress changes to the lattice structure. Wang et al.[160] prepared a NVPF composite [Na3.3K0.2V1.5Mg0.5(PO4)2F3] with N-doped carbon coating, employing a binary co-doping strategy. The surface of the N-doped carbon layer was rich in defects, providing additional active sites and enhancing conductivity. Such modifications accelerate the reaction kinetics and augment electrolyte interaction through an expanded specific surface area, thus streamlining the electrochemical process. Concurrently, strategic heteroatom substitution leads to a more efficient engagement of Na+ in the electrochemical activities, thereby bolstering the cathode’s structural integrity. The ionic radius of Bi3+ is larger than that of V, which is beneficial for Na+ intercalation by increasing the interplanar spacing. A Bi-doped carbon-coated NVPF [Na3V2-xBix(PO4)2F3/C] was synthesized using the sol-gel method combined with high-temperature calcination[161]. The results indicate that the incorporation of an optimal amount of Bi-doping and carbon coating significantly enhanced the conductivity, Na+ diffusion performance, and structural stability of the cathode material during the charge-discharge cycle.

Replacing V3+ with Ca2+ can reduce the bandgap, enhance electronic conductivity, and improve structural stability. Puspitasari et al.[162] prepared a Ca-doped NVPF cathode material with a carbon coating [Ca-doped Na3V2(PO4)2F3 with a carbon coating (NVPFCa-x/C)] using sol-gel and carbothermal reduction techniques. The results have shown that Ca2+ doping did not alter the crystal structure but expanded the lattice spacing. The incorporation of the Ca2+ ion leads to the emergence of two localized states within the band gap, thereby significantly improving the electrochemical performance in comparison to Mg-doped NVPF/C.

Doping trivalent rare-earth ions at the V3+ site is an effective strategy for improving NVPF performance. In a recent study, an atomic simulation method based on the minimization of lattice energy was used to explore the effects of rare-earth ion doping on the local structural deformation and intrinsic defect properties of the NVPF cathode material[163]. The study found that defects formed by Na interstitials and vacancies (Na Frenkel defects) serve to promote Na+ diffusion. In addition, the concentration of Na vacancies and interstitials can be increased by doping rare-earth ions at Na and P sites, enhancing reaction kinetics.

The removal or insertion of excess Na+ in NVPF can be achieved within a relatively wide voltage range by adding Al. A study of Al-doped NVPF (NVAlPF-x) synthesis has demonstrated that Al doping reduced the band gap without causing a phase transition[15]. Doping with Al increased the Na+ diffusion coefficient, and Al3+ successfully replaced the V3+ site within the NVPF lattice.

A high-entropy doping can suppress adverse phase transitions at low voltages, promoting Na+ storage at high voltages. Gu et al.[60] have applied the concept of high entropy with respect to the PACs lattice, and prepared a carbon-free high-entropy Na3V1.9(Ca,Mg,Al,Cr,Ca)0.1(PO4)2F3 cathode material by partially substituting V atoms. Density functional theory (DFT) calculations have shown that high-entropy doping alters the electronic state, enhancing conductivity and Na+ migration rate, further elevating the average voltage platform to 3.81 V and increasing the energy density to 445.42 Wh kg-1.

During the electrode reaction, new active redox pairs can be formed by high-activity elements, such as Mn2+[164], Cr3+[118], and Ti3+[63], which serve to appreciably enhance the specific capacity and output voltage[165]. Chromium doping is particularly effective as Cr3+ has a similar ionic radius to V3+. A moderate level of Cr3+ doping can introduce crystal defects, reducing the Na+ transport distance, and enhancing diffusion with higher electron transfer and charge transfer rates. For instance, Cr-doped NVPF composites with a carbon coating were synthesized using Cr3+ as an iso-valent doping ion by the sol-gel method[166] [Figure 5B]. The results established improved electrical conductivity as a result of lower charge transfer resistance (Rct) due to Cr3+ doping. In addition, Yi et al.[167] prepared Cr-doped NVPF composites [Na3V2-xCrx(PO4)2F3/C] by sol-gel. The enhanced electrochemical performance is attributed to the elevated crystallinity, optimized intrinsic electronic conductivity, and Na+ diffusion coefficient resulting from the doping of Cr3+. This doping accelerates the electron transport and charge transfer rate.

An appropriate level of Mn2+ incorporation in NVPF can effectively enhance the Na+ diffusion rate due to the larger radius of Mn2+ relative to V3+ that regulates the lattice and surface energy. Gu et al.[168] synthesized NVPF-NS with a 2D NS structure, which enabled an adjustment of the surface energy and V site in NVPF by partially replacing V3+ with Mn2+. This led to combined presence of V3+ and V4+, which enhanced NVPF electronic conductivity and structural stability. NVPF materials doped with different valence Mn ions have been produced in a hydrothermal procedure[169]. The results reported demonstrated that Mn doping altered the NVPF crystal structure and residual carbon content [Figure 5C]. The incorporation of Mn2+ resulted in a NVPF micro-rectangular morphology, restricting crystal size, reducing Rct and promoting Na+ diffusion.

Doping with Ti can effectively inhibit the formation of an irreversible tetragonal phase, and achieve Na+ storage via a reversible transition from the Amam to Cmc21 phase. Tong et al.[170] synthesized Ti-doped NVPF (NVTPF) cathode material by solid-state sintering [Figure 5D]. The incorporation of Ti stabilized the local structure via F-Ti bond formation by precipitation, with minimal changes during charge and discharge. In the electrochemical process, vanadium was reversibly transformed from V3+ to V5+, where Ti4+ was unchanged. The substitution of W6+ at the V site can improve Na+ diffusion and electronic conductivity, while inhibiting grain growth. W-doped carbon-coated NVPF synthesis using a sol-gel and ball-milling assisted by sonication has been shown to enhance structural integrity and improve Na+ mobility[171].

The element-doped NVPF with excellent rate performance reported in recent studies is illustrated in Figure 5E-G. A comparison of the electrochemical performance of NVPF cathodes in SIBs prepared using different doping strategies is presented in Table 2. It is evident that ion doping can enhance the intrinsic conductivity of NVPF, maintaining the stability of the crystal structure. Doping with inexpensive metals can reduce production costs. However, the incorporation of ions with large radii and excessive doping may damage the crystal structure and lower material stability. In order to ensure effective doping, many factors such as element type, doping amount, distribution uniformity, dopant siting and cost require further comprehensive investigation.

Table 2

Comparison of the electrochemical performance of NVPF cathodes in SIBs prepared using different elemental doping strategies

CompositeDoped elementElectrolyteElectrochemical performances
(current density, discharge capacity, cycles, and capacity retention)
Ref.
N1-xKxVPF/CK1 M NaClO4 in EC: PC (1:1) with 5 vol.% FEC0.2 C, 128 mAh g-1, 100, 92% (0.2 C)[153]
NVPF-HE@Chigh-entropy elements (Mg, Cr, Al, Nb)1 M NaClO4 in EC/PC1 C, 128 mAh g-1, 800, 83% (5 C)[154]
Fe-NVPF@N-CNTsFe1 M NaClO4 in DEC:EC = 1:1 vol% with 5 % FEC0.1 C, 105 mAh g-1, 1,200, 83.38% (5 C)[155]
Na3.3K0.2V1.5Mg0.5(PO4)2F3K, Mg1 M NaClO4 in EC: DMC: DEC (1:1:1, vol)1 C, 115.6 mAh g-1, 1,000, 93.1% (10 C)[160]
Na3V2-xBix(PO4)2F3/CBi1 M NaClO4 in EC: DEC (1:1) with 5 vol.% FEC0.1 A g-1, 107.4 mAh g-1, 100, 90.4% (0.1 A g-1)[161]
NVPFCa-x/CCa1 M NaClO4 in EC: PC (1:1, by vol) 0.1 C, 124 mAh g-1, 1,000, 70% (10 C)[162]
NVAlPF-xAl1 M NaClO4 in EC: DEC (1:1) with 5 vol.% FEC0.1 C, 121.3 mAh g-1, 400, 75% (5 C)[15]
Na3V2-xCrx(PO4)2F3/CCr1 M NaClO4 in PC with 5 vol. % FEC0.2 C, 117 mAh g-1, 1,000, 68.7% (10 C)[167]
NVPF-NSMn1 M NaClO4 in PC with 5 vol. % FEC0.1 C, 120 mAh g-1, 1,000, 80.7% (1 C)[168]
NVMPF@CMn1 M NaClO4 in EC: DMC (1:1) with 5 vol.% FEC0.2 C, 116.2 mAh g-1, 400, 67.7% (1 C)[169]
NVTPFTiNaPF6 in PC/EC/DMC (1:1:1, by vol)0.1 C, 70 mAh g-1, 100, 90% (0.1 C)[170]
NVPFWW1 M NaClO4 in EC: PC (1:1, by vol) with 3 vol.% FEC0.1 C, 140 mAh g-1, 250, 97.4% (2 C)[171]

OPTIMIZATION OF SIBS

Researchers have employed a number of strategies to enhance the electrochemical performance of NVPF in SIBs, such as interface and defect engineering [172], adjustment of the electrode structure[173,174], and optimization of electrolyte composition[175-177]. The associated mechanisms that underpin electrochemical performance have also been investigated.

Interface and defect engineering

In recent years, researchers have made numerous attempts to improve NVPF electrochemical performance by utilizing interface and defect engineering to enhance electronic transmission dynamics and ionic diffusion kinetics, mitigating the structural stress generated during electrochemical processes.

Defect engineering can optimize Na+ transport kinetics. By introducing ordered vacancies, Na+ can be uniformly adsorbed, alleviating mechanical stress during charging and discharging and significantly improving rate performance[178]. Bulk nitrogen doping in NVPF can expand the interlayer spacing and enhance electronic conductivity, raising the voltage platform while achieving ultra-long cycle stability[179]. Vacancies or doping defects can lower the migration energy barrier for sodium ions and increase the diffusion coefficient.

Interfacial engineering can enhance structural stability and interface compatibility. The construction of strong chemical bond interfaces can inhibit surface side-reactions and improve electron transport efficiency[180]. Hard carbons are the most commonly used anode materials in SIBs, while they generally suffer from poor cycling and rate capability. Glyme-ether-based electrolytes enable long-term cycling of SIBs based on hard carbon anode and NVPF cathode, due to the formation of fluorine-rich SEI layer with superior stability[181]. Applying carbon coating or compositing with conductive materials (such as graphene) reduces interface impedance and alleviates volume expansion, improving the cycling stability of NVPF[103,108,128,145,150].

Interface and defect engineering can significantly improve the rate performance, cycle life, and low-temperature adaptability of NVPF cathodes by regulating Na+ transport pathways, enhancing electronic conductivity, and optimizing interface stability. Future research should focus on a precise design of defects and interfaces, with an examination of synergistic optimization mechanisms.

Optimization of electrolyte composition

The electrolyte provides the necessary medium for ionic transfer between cathode and anode[182]. There are currently two main approaches to optimizing SIB electrolytes. The first deals with ionic conductivity and viscosity, focusing on the type or composition of the electrolyte. The second considers the use of electrolyte additives to form a stable protective layer at the interface between the electrolyte and the electrode, enhancing battery cycling performance.

The operation of SIBs at temperatures above 50 °C presents challenges in terms of safety and stable cycling, requiring high-temperature-resistant liquid electrolytes. Liang et al.[183] combined fluoroethylene carbonate (FEC) and trimethylsilyl phosphite (TMSPi) to prepare a non-flammable electrolyte. This dual-functional integrated strategy resulted in a non-combustible electrolyte with thermal stability that ensured safe and stable battery operation at high temperatures [Figure 6A and B]. In addition, the same research group considered a film-forming modulation strategy, which improved the ionic transport capability and structural/thermal stability of CEI/SEI (solid electrolyte interphase) by optimizing the interfacial formation of P/Si intermediates, enabling stable operation of the SIB over a wide temperature range (-25 °C to 75 °C), successfully achieving all-weather adaptability [Figure 6C][184]. The assembled HC//NVPF full cell maintained over 90% capacity retention following long-term testing at -25 °C and 50 °C. The addition of an organic co-solvent can improve the low temperature resistance of the battery. A fully-printed flexible aqueous rechargeable SIB (ARSIB) with superior mechanical properties was assembled using screen printing technology for the first time, with NVPF/C as the cathode material and Na3MnTi(PO4)3/C as the anode material[185]. A 17 m (mol kg-1) NaClO4-EG mixed electrolyte which can withstand low temperature (-20 °C) was prepared by incorporating ethylene glycol (EG) as co-solvent with the 17 m NaClO4 electrolyte. This flexible battery had a capacity of 40.1 mAh g-1, with outstanding mechanical and cycling performance [Figure 6D and E].

Recent research progress in the application of Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>2</sub>F<sub>3</sub> cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization

Figure 6. (A) The TMSPi-induced film formation process; (B) Flame test carried out with separators soaked in different electrolytes. Reproduced with permission[183]. Copyright 2024, Elsevier; (C) Galvanostatic charge-discharge (GCD) curves associated with continuous temperature testing simulating a wide range of all-weather conditions. Reproduced with permission[184]. Copyright 2024, American Chemical Society; (D) Power to electronic devices at 25 °C and -20 °C; (E) GCD curves of a battery with different bending angles at 0.1 C. Reproduced with permission[185]. Copyright 2024, Wiley; (F) V dissolution of a charged NVPF powder in EC-PC-DMC with different sodium salts at 55 °C for 21 days. Reproduced with permission[186]. Copyright 2023, Elsevier; (G) Schematic representation of solvation structure; (H) TEM image of a circulating NVPF cathode in NPP6 electrolyte (PFPN as a co-solvent in electrolyte at molar ratio of PC:PFPN = 7.5:1). Reproduced with permission[188]. Copyright 2024, Wiley -VCH. TMSPi: trimethylsilyl phosphite; NVPF: Na3V2(PO4)2F3; EC: ethylene carbonate; PC: propylene carbonate; DMC: dimethyl carbonate; TEM: transmission electron microscope; NPP6: 6th sodium-ion electrolyte formulation with PC and PFPN; PFPN: ethoxy (pentafluoro) cyclotriphosphazene.

The dissolution of V and resultant parasitic reactions are widely recognized as key factors leading to declining battery performance. A combined use of surface coatings and changes in electrolyte formulation was proposed to reduce V dissolution [Figure 6F][186]. It was shown that electrolytes based on sodium trifluoromethane sulfonimide (NaTFSI) inhibited V dissolution, but degradation must be prevented at high potentials. In addition, a carbon coating of NVPF particles with the addition of electrolyte surface protective agents is effective in minimizing V dissolution.

The formation of an anion-derived CEI layer improves interfacial stability and promotes rapid transfer of Na+ at the interface. The NVPF cathode exhibited greater stability in the diglyme-based electrolyte. Zheng et al.[187] have reported a thorough examination of the Na storage mechanism by NVPF in diglyme and propylene carbonate(PC)/ethylene carbonate (EC) electrolytes containing 0.5 M NaPF6 at low temperatures. This electrolyte system significantly enhanced the rate of Na+ diffusion in the bulk electrolyte with a reduction in the charge transfer barrier to 158.6 meV, which is only one-third of that in a PC/EC-based electrolyte. Moreover, the battery can function normally and maintain stable performance at -60 °C.

Recently, a novel ternary low-concentration electrolyte and cation/anion solvation strategy was proposed, which utilized ethoxy (pentafluoro) phosphazene (PFPN) as a weakly polar solvent to improve the electrode/electrolyte interfacial stability and enhance ion conductivity and oxidation stability [Figure 6G and H][188]. This strategy enabled increased cycling performance and energy density, facilitating durable high-voltage SIBs. The ionic conductivity of the low-concentration electrolyte was increased to 6.12 mS cm-1, and the electrochemical stability window was raised to 4.84 V.

The fundamental cause of the significant capacity decline of NVPF3-based batteries is associated with the continuous electrolyte oxidation process rather than F loss or the collapse of the crystal structure. The capacity retention of NVPF3 batteries can be improved by optimizing the electrolyte. Jiang et al.[189] designed an electrolyte additive containing two -C≡N groups that enabled a stable SIB cycling at high voltage.

Sulfonate anion-based ionic liquids have been widely used as electrolytes for secondary batteries. An inhibitory effect of the Na[FSA]-[C2C1im][FSA] ionic liquid on Al corrosion in the presence of [FAP]- has been reported[190]. A uniform deposition passivation layer formed by the decomposition products of [FSA]- and [FAP]- mitigated the corrosion of Al after adding [FAP]-. The study found that Na[FSA]-[C2C1im][FSA] moderately inhibited Al corrosion at 25 °C, but a severe pitting corrosion of the Al electrode was observed at 90 °C when the potential was higher than 4 V vs Na+/Na. In the charge-discharge tests of NVPF electrodes, the addition of [FAP]- elevated the electrochemical performance of the ionic liquid electrolyte, particularly at high temperatures.

Exogenous additives

Strategies for increasing the exposure of active crystal facets that promote Na+ transport involving selective crystal growth are crucial to developing cathodes for high-performance SIBs[191]. For example, NVPFs with different (002) face exposure rates were synthesized using carbon clusters (c-clusters) as inducers to regulate the surface energy [Figure 7A][192]. The study demonstrated that the (002) face possessed more stable Na+ storage sites and lower diffusion barriers, where the adsorption of c-clusters generated 2D particles that influenced the crystal growth direction. The optimized NVPFs exhibited a high reversible capacity (123 mAh g-1) and energy density (431 Wh kg-1).

Recent research progress in the application of Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>2</sub>F<sub>3</sub> cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization

Figure 7. (A) Schematic representation of NVPF growth. Reproduced with permission[192]. Copyright 2024, Royal Society of Chemistry; (B) Electron cloud density of associated materials. Reproduced with permission[193]. Copyright 2024, Wiley -VCH; (C) TEM image of NVPF particle; (D) Schematic representation of the effects of NaPAA and PVDF binders on the formation of CEI layers. Reproduced with permission[194]. Copyright 2024, Royal Society of Chemistry; (E) Schematic representation of the Na4C6O6 loss compensation mechanism for the cathode and anode. Reproduced with permission[195]. Copyright 2022, Royal Society of Chemistry. NVPF: Na3V2(PO4)2F3; TEM: transmission electron microscope; NaPAA: sodium polyacrylate; PVDF: polyvinylidene fluoride;‌ CEI: cathode-electrolyte interface.

Chemical pre-precipitation (CP) is an effective method to enhance SIB energy density. Current sodium agents are mostly polycyclic aromatic hydrocarbons with low reduction potential, which can readily result in excessive sodium insertion in the cathode with structural deterioration. As novel CP agents, aromatic ketones can endow the agents with mild reducibility by introducing the carbonyl (C=O) function to balance conjugation and inductive effects [Figure 7B][193]. While maintaining structural stability, the charge capacity of the positive electrode material was significantly improved.

Introducing a sodium polyacrylate (NaPAA) binder to the NVPF cathode can overcome the instability of the interface[194]. The NaPAA binder interacted with the NVPF cathode via ionic-dipole interactions, improving the distribution of active materials and forming a stable ionic conductive NaPO2F2-rich CEI layer that effectively inhibited electrolyte decomposition [Figure 7C and D]. Zhang et al.[195] prepared a novel high-capacity self-sacrificial Na4C6O6 additive, which contained a large number of sulfonic carboxyl groups [Figure 7E]; the resultant capacity exceeded 400 mAh g-1. This additive can provide sufficient Na to compensate for Na loss during the initial charge-discharge cycle. The energy density of the full cell with 9% additive (HC//NVPF/rGO) increased from 154.5 Wh kg-1 to 210.8 Wh kg-1.

Mechanism of SIB performance

As the battery interphase properties and chemical composition vary with temperature and the dissolution and recombination of the interphase are affected by temperature, the desired characteristics can be achieved by combining high and low temperatures. A recent study has illustrated the effect of temperature on the interphase properties of Na3V2(PO4)2F3|hard carbon (NVPF|HC) batteries [Figure 8A][196]. The results demonstrated that cycle stability, impedance, and rate performance were sensitive to temperature, leading to different SEI formation pathways. The battery capacity retention and coulomb efficiency values were lowest at 25 °C. At higher temperatures, the battery was stable but exhibited a high impedance, whereas the NaF content was high with minimal transition metal dissolution at low temperatures. Superior reversible capacity, cycle stability, and high-power rate capability can be achieved by adjusting temperature.

Recent research progress in the application of Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>2</sub>F<sub>3</sub> cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization

Figure 8. (A) Effect of temperature on different NVPF|HC cells parameters. Reproduced with permission[196]. Copyright 2023, IOP Publishing; (B) Performance comparison at different temperatures. Reproduced with permission[197]. Copyright 2023, American Chemical Society; (C) Rate capacity of the batteries; (D) Nyquist plots for NVPF-HU and NVPF-UU; (E) Effect of heat treatment on the failed NVPF-H cathode. Reproduced with permission[200]. Copyright 2022, Wiley -VCH. NVPF|HC: Na3V2(PO4)2F3|hard carbon.

Komayko et al.[197] investigated the effects of two different intercalation approaches on the performance of SIB cathode materials at low temperatures and high discharge current limits [Figure 8B]. The research established that NaVPO4F, which followed a solid solution mechanism, was significantly superior to NVPF which exhibited phase separation during Na+ intercalation. The energy density retention of NaVPO4F at 10 C and -30 °C was 387 and 308 Wh kg-1, respectively, compared with 259 and 184 Wh kg-1 for NVPF. Although the Na+ diffusion coefficient in the single-phase region was similar for both materials, the decline in performance from the quasi-equilibrium charge-discharge state was caused by concentration polarization effects. In the case of NVPF, the slow movement of phase boundaries during phase transformation was the principal cause of the observed rapid capacity decay.

Molecular dynamics studies of the local nanostructure associated with a series of carbonate-based sodium electrolytes at the NVP and NVPF interfaces were conducted with an analysis of de-solvation[198]. The results have indicated that the solvent filler behavior at the electrode surface is a key factor affecting the energy barrier height related to de-solvation. On a graphite electrode without SEI, no dissolution was observed due to the absence of specific oriented interactions between the solvent and the carbon sheet. However, in polyanion nested compounds with uneven surfaces and donor and acceptor groups, the closest layer of solvent molecules exhibited a distinct arrangement, with a partial loss of the Na+ solvation shell on passage through the layer. This effect was particularly pronounced on the NVP interface, and the high energy barrier was attributed to the highly ordered and well-oriented first solvent layer. The arrangement of the solvent was less pronounced at the NVPF interface due to the presence of non-polar fluorine atoms which interfered with the electrostatic attraction between the solvent and electrode.

Semykina et al.[199] studied the mechanism and kinetics of the solid-phase reaction between NaF and VPO4, along with the effect of high-energy ball-milling (HEBM). The results showed that the reaction proceeded in a “dimensionality reduction” manner, where the rate-limiting step involved the diffusion of Na+ and F- through the single VPO4 side. The macroscopic HEBM mechanism followed a third-order reaction model, enabling a significant increase in the rate constant and diffusion coefficient. The apparent activation energy was approximately 290 kJ mol-1. The similarity in the structure of reactant (VPO4) and product (NVPF) facilitated interaction with a partial occupation of Na+ sites that favored Na+ migration. The electrochemical performance of the prepared NVPF matched conventional materials.

An analysis of the interface between a high-voltage NVPF cathode and electrolyte in SIBs has revealed that the interface phase of high-concentration electrolytes was primarily composed of anions, including a significant presence of anion derivatives (CxClOy) and inorganic substances (Na2CO3 and NaF), which increased resistance to Na+ transport and exacerbated gas evolution[200]. In low-concentration electrolytes, the solvation structure of the contact ion pairs was conducive to forming a stable anion-regulated and solvent-derived composite interface phase, enhancing the cycling stability and the interfacial kinetic performance of the battery at 55 °C [Figure 8C and D]. In addition, the performance of the failed NVPF cathode could be restored by annealing [Figure 8E]. The CxClOy phase decomposed in argon at 300 °C, reconstructing the interface phase.

SUMMARY AND OUTLOOK

SIBs, exhibiting unique advantages with respect to the cost-effective supply of sodium and high-power density, are considered viable alternatives to LIBs. The use of NVPF with a NASICON structure as cathode material offers high operating potential and structural stability that has attracted significant research activity in SIB development and applications. This review has addressed the research progress made in the past three years regarding NVPF cathode materials for use in SIBs, examining synthesis methods, modification strategies, and component optimization. Despite significant advancements in NVPF development, numerous challenges lie ahead. We have summarized the technological progress and challenges [Figure 9], and anticipate the following future directions for the optimization of high-performance NVPF cathodes:

Recent research progress in the application of Na<sub>3</sub>V<sub>2</sub>(PO<sub>4</sub>)<sub>2</sub>F<sub>3</sub> cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization

Figure 9. The progress in technology and future challenges of NVPF in SIBs. NVPF: Na3V2(PO4)2F3; SIB: sodium-ion battery; SEI: solid electrolyte interphase; CEI: cathode-electrolyte interface.

(1) Synthesis of nano-structured NVPF can optimize Na+ diffusion and enhance the interfacial contact between the active material and electrolyte. Existing studies of NVP morphology are insufficient, and there is an urgent need for a better understanding of the factors that control morphology. Furthermore, integrating multi-dimensional carbon materials to construct high-quality hierarchical carbon layers, forming abundant electron and ion transport pathways, represents a critical research direction. An exploration of other coating materials, such as Al2O3 and RuO2[201], can further develop effective composite synthesis strategies.

(2) Current research on modification mechanisms is insufficient, particularly concerning theoretical models of NVPF crystal growth and an in-depth exploration of doping mechanisms. The chemical synergism associated with the crystalline NVPF positive electrode materials and the mechanism of valence changes may be revealed by utilizing advanced characterization techniques, such as solid-state nuclear magnetic resonance, X-ray absorption spectroscopy, and combining these measurements with first-principles calculations based on DFT[94].

(3) It is particularly important to develop effective Na+ compensation methods to address the issue of reduced energy density due to Na loss. Introducing suitable additives to NVPF or the electrolyte represents a feasible approach. It is crucial to thoroughly investigate the effects of electrolyte formulations, concentrations, and types of additives, focusing on such parameters as coulomb efficiency, cycle life, and rate performance[182]. It is essential to develop electrolytes that are stable over a wider temperature range in order to satisfy the requirements of practical application[202]. The research and development of solid-state electrolytes represent a potential means of enhancing the electrochemical performance of NVPF-based batteries at high temperatures.

(4) The role of binders must be considered, with systematic testing in order to screen and identify the most suitable binder. This must include an assessment of the physical and chemical properties of the binder in relation to NVPF, and the associated impact on performance[194]. In addition, the production of structures without the need for binders should be examined.

This review has offered new insights into the synthesis and design of high-performance NVPF cathode materials. Given the rapid progress of SIB technology, we anticipate the emergence of a series of NVPF cathode materials with superior electrochemical properties in the near future, paving the way for commercial applications.

DECLARATIONS

Acknowledgments

We thank EditSprings (https://www.editsprings.cn) for the expert linguistic assistance during the preparation of this manuscript.

Authors’ contributions

Conceptualization: Liang, Y.; Chen, Z.; Sun, Z.

Writing-original draft: Liang, Y.

Writing of review and editing: Chen, Z, Sun, Z.

Supervision: Ning, D.; Chen, Z.; Sun, Z.

Project administration: Xu, L.; Chen, Z.; Sun, Z.

Availability of data and materials

Not applicable.

Financial support and sponsorship

The authors acknowledge the National Natural Science Foundation of China (Nos. 52202304, 22473032), Guangdong Province Science and Technology Plan Project (Nos. 2024B1212050006, 2022A0505020010), Guangdong Provincial University Key Project (No. 2022ZDZX3043), Zhuhai Innovation and Entrepreneurship Team Project (No. 22055905200017), Zhuhai Science and Technology Department Project (Nos. 2420004000114, 2420004000003, 2320004000233), Guangdong University of Technology Hundred Talents Program (No. 263118136), and Marine Biomaterials Laboratory of ZIAT and Dangan Town.

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2025.

REFERENCES

1. Obama, B. The irreversible momentum of clean energy. Science 2017, 355, 126-9.

2. Deng, X.; Li, J.; Ma, L.; Sha, J.; Zhao, N. Three-dimensional porous carbon materials and their composites as electrodes for electrochemical energy storage systems. Mater. Chem. Front. 2019, 3, 2221-45.

3. Zeng, X.; Xu, Y.; Yin, Y.; Wu, X.; Yue, J.; Guo, Y. Recent advances in nanostructured electrode-electrolyte design for safe and next-generation electrochemical energy storage. Mater. Today. Nano. 2019, 8, 100057.

4. Liu, X.; Wang, Y.; Yang, Y.; et al. A MoS2/Carbon hybrid anode for high-performance Li-ion batteries at low temperature. Nano. Energy. 2020, 70, 104550.

5. Usiskin, R.; Lu, Y.; Popovic, J.; et al. Fundamentals, status and promise of sodium-based batteries. Nat. Rev. Mater. 2021, 6, 1020-35.

6. Perveen, T.; Siddiq, M.; Shahzad, N.; Ihsan, R.; Ahmad, A.; Shahzad, M. I. Prospects in anode materials for sodium ion batteries - a review. Renew. Sust. Energ. Rev. 2020, 119, 109549.

7. Fan, E.; Li, L.; Wang, Z.; et al. Sustainable recycling technology for Li-Ion batteries and beyond: challenges and future prospects. Chem. Rev. 2020, 120, 7020-63.

8. Jiao, J.; Wu, K.; Li, N.; et al. Tuning anionic redox activity to boost high-performance sodium-storage in low-cost Na0.67Fe0.5Mn0.5O2 cathode. J. Energy. Chem. 2022, 73, 214-22.

9. Qi, Y.; Tong, Z.; Zhao, J.; et al. Scalable room-temperature synthesis of multi-shelled Na3(VOPO4)2F microsphere cathodes. Joule 2018, 2, 2348-63.

10. Hwang, J. Y.; Myung, S. T.; Sun, Y. K. Sodium-ion batteries: present and future. Chem. Soc. Rev. 2017, 46, 3529-614.

11. Rajagopalan, R.; Tang, Y.; Jia, C.; Ji, X.; Wang, H. Understanding the sodium storage mechanisms of organic electrodes in sodium ion batteries: issues and solutions. Energy. Environ. Sci. 2020, 13, 1568-92.

12. Xie, F.; Chen, J.; Xu, J.; et al. Investigation of Na7V3(P2O7)4/carbon composite as cathode material for sodium-ion battery: influences of carbon additives and voltage windows. J. Energy. Storage. 2024, 77, 109855.

13. Fang, Y.; Zhang, J.; Xiao, L.; Ai, X.; Cao, Y.; Yang, H. Phosphate framework electrode materials for sodium ion batteries. Adv. Sci. (Weinh). 2017, 4, 1600392.

14. Chen, S.; Wu, C.; Shen, L.; et al. Challenges and perspectives for NASICON-type electrode materials for advanced sodium-ion batteries. Adv. Mater. 2017, 29.

15. Zhuang, S.; Yang, C.; Zheng, M.; et al. A combined first principles and experimental study on Al-doped Na3V2(PO4)2F3 cathode for rechargeable Na batteries. Surf. Coat. Tech. 2022, 434, 128184.

16. Zhou, L.; Zhuang, Z.; Zhao, H.; Lin, M.; Zhao, D.; Mai, L. Intricate hollow structures: controlled synthesis and applications in energy storage and conversion. Adv. Mater. 2017, 29.

17. Sun, D.; Ye, D.; Liu, P.; et al. MoS2/graphene nanosheets from commercial bulky MoS2 and graphite as anode materials for high rate sodium-ion batteries. Adv. Energy. Mater. 2018, 8, 1702383.

18. Wang, W.; Li, W.; Wang, S.; Miao, Z.; Liu, H. K.; Chou, S. Structural design of anode materials for sodium-ion batteries. J. Mater. Chem. A. 2018, 6, 6183-205.

19. Xue, X.; Sun, D.; Zeng, X.; et al. Two-step carbon modification of NaTi2(PO4)3 with improved sodium storage performance for Na-ion batteries. J. Cent. South. Univ. 2018, 25, 2320-31.

20. Huang, Z.; Hou, H.; Wang, C.; Li, S.; Zhang, Y.; Ji, X. Molybdenum phosphide: a conversion-type anode for ultralong-life sodium-ion batteries. Chem. Mater. 2017, 29, 7313-22.

21. Sun, D.; Luo, B.; Wang, H.; Tang, Y.; Ji, X.; Wang, L. Engineering the trap effect of residual oxygen atoms and defects in hard carbon anode towards high initial Coulombic efficiency. Nano. Energy. 2019, 64, 103937.

22. Liu, N.; He, Z.; Zhu, J.; et al. Crystal doping of K ion on Na site raises the electrochemical performance of NaTi2(PO4)3/C anode for sodium-ion battery. Ionics 2020, 26, 3387-94.

23. Lao, M.; Zhang, Y.; Luo, W.; Yan, Q.; Sun, W.; Dou, S. X. Alloy-based anode materials toward advanced sodium-ion batteries. Adv. Mater. 2017, 29.

24. Hao, Y.; Li, X.; Liu, W.; et al. Depolarization of Li-rich Mn-based oxide via electrochemically active Prussian blue interface providing superior rate capability. Carbon. Energy. 2023, 5, e272.

25. Su, H.; Jaffer, S.; Yu, H. Transition metal oxides for sodium-ion batteries. Energy. Storage. Mater. 2016, 5, 116-31.

26. Cao, Y.; He, Y.; Gang, H.; et al. Stability study of transition metal oxide electrode materials. J. Power. Sources. 2023, 560, 232710.

27. Sang, L.; Wang, K.; Wu, Y.; Ma, C. The improved solar weighted absorptance and thermal stability of desert sand coated with transition metal oxides for direct particle receiver. Sol. Energy. Mater. Sol. Cells. 2023, 251, 112158.

28. Rao, T.; Zhou, Y.; Jiang, J.; Yang, P.; Liao, W. Low dimensional transition metal oxide towards advanced electrochromic devices. Nano. Energy. 2022, 100, 107479.

29. Liu, Z.; Huang, Y.; Huang, Y.; et al. Voltage issue of aqueous rechargeable metal-ion batteries. Chem. Soc. Rev. 2020, 49, 180-232.

30. Camacho, P. S.; Wernert, R.; Duttine, M.; et al. Impact of synthesis conditions in Na-rich Prussian blue analogues. ACS. Appl. Mater. Interfaces. 2021, 13, 42682-92.

31. Li, J.; Yi, H.; Xiao, Y.; et al. Freestanding catalytic membranes assembled from blade-shaped Prussian blue analog sheets for flow-through degradation of antibiotic pollutants. Appl. Catal. B:. Environ. 2023, 336, 122922.

32. Chun, J.; Wang, X.; Wei, C.; Wang, Z.; Zhang, Y.; Feng, J. Flexible and free-supporting Prussian blue analogs /MXene film for high-performance sodium-ion batteries. J. Power. Sources. 2023, 576, 233165.

33. Han, J.; Hu, Y.; Han, Q.; Liu, X.; Wang, C. Synthesis of high-specific-capacity Prussian blue analogues for sodium-ion batteries boosted by grooved structure. J. Alloys. Compd. 2023, 950, 169928.

34. Xu, N.; Lei, H.; Hou, T.; et al. Constructing an asymmetric supercapacitor based on Prussian blue analogues-derived cobalt selenide nanoframeworks and iron oxide nanoparticles. Electrochimica. Acta. 2023, 439, 141686.

35. Yang, Y.; Guo, C.; Zeng, Y.; Luo, Y.; Xu, J.; Wang, C. Peroxymonosulfate activation by CuFe-Prussian blue analogues for the degradation of bisphenol S: effect, mechanism, and pathway. Chemosphere 2023, 331, 138748.

36. Okubo, M.; Li, C. H.; Talham, D. R. High rate sodium ion insertion into core-shell nanoparticles of Prussian blue analogues. Chem. Commun. (Camb). 2014, 50, 1353-5.

37. He, M.; Davis, R.; Chartouni, D.; et al. Assessment of the first commercial Prussian blue based sodium-ion battery. J. Power. Sources. 2022, 548, 232036.

38. Yue, Y.; Binder, A. J.; Guo, B.; et al. Mesoporous Prussian blue analogues: template-free synthesis and sodium-ion battery applications. Angew. Chem. Int. Ed. Engl. 2014, 53, 3134-7.

39. Peng, J.; Zhang, W.; Liu, Q.; et al. Prussian blue analogues for sodium-ion batteries: past, present, and future. Adv. Mater. 2022, 34, e2108384.

40. Wang, W.; Gang, Y.; Hu, Z.; et al. Reversible structural evolution of sodium-rich rhombohedral Prussian blue for sodium-ion batteries. Nat. Commun. 2020, 11, 980.

41. Dong, Y.; Di, S.; Zhang, F.; et al. Nonaqueous electrolyte with dual-cations for high-voltage and long-life zinc batteries. J. Mater. Chem. A. 2020, 8, 3252-61.

42. Ni, Q.; Bai, Y.; Wu, F.; Wu, C. Polyanion-type electrode materials for sodium-ion batteries. Adv. Sci. (Weinh). 2017, 4, 1600275.

43. Zhao, L.; Zhang, T.; Zhao, H.; Hou, Y. Polyanion-type electrode materials for advanced sodium-ion batteries. Mater. Today. Nano. 2020, 10, 100072.

44. Yuan, Y.; Wei, Q.; Yang, S.; et al. Towards high-performance phosphate-based polyanion-type materials for sodium-ion batteries. Energy. Storage. Mater. 2022, 50, 760-82.

45. Gao, Y.; Zhang, H.; Liu, X.; et al. Low-cost polyanion-type sulfate cathode for sodium-ion battery. Adv. Energy. Mater. 2021, 11, 2101751.

46. Yuan, D.; Liang, X.; Wu, L.; et al. A honeycomb-layered Na3Ni2SbO6: a high-rate and cycle-stable cathode for sodium-ion batteries. Adv. Mater. 2014, 26, 6301-6.

47. You, Y.; Wu, X.; Yin, Y.; Guo, Y. High-quality Prussian blue crystals as superior cathode materials for room-temperature sodium-ion batteries. Energy. Environ. Sci. 2014, 7, 1643-7.

48. Wang, L.; Song, J.; Qiao, R.; et al. Rhombohedral Prussian white as cathode for rechargeable sodium-ion batteries. J. Am. Chem. Soc. 2015, 137, 2548-54.

49. Jin, T.; Li, H.; Zhu, K.; Wang, P. F.; Liu, P.; Jiao, L. Polyanion-type cathode materials for sodium-ion batteries. Chem. Soc. Rev. 2020, 49, 2342-77.

50. Xiang, X.; Zhang, K.; Chen, J. Recent advances and prospects of cathode materials for sodium-ion batteries. Adv. Mater. 2015, 27, 5343-64.

51. Liang, K.; Wu, D.; Ren, Y.; Huang, X.; Ma, J. Research progress on Na3V2(PO4)2F3-based cathode materials for sodium-ion batteries. Chin. Chem. Lett. 2023, 34, 107978.

52. Rajagopalan, R.; Tang, Y.; Ji, X.; Jia, C.; Wang, H. Advancements and challenges in potassium ion batteries: a comprehensive review. Adv. Funct. Mater. 2020, 30, 1909486.

53. Chen, Y.; Zhao, Y.; Tian, S.; Wang, P.; Qiu, F.; Yi, T. Recent progress and strategic perspectives of high-voltage Na3V2(PO4)2F3 cathode: fundamentals, modifications, and applications in sodium-ion batteries. Compos. B:. Eng. 2023, 266, 111030.

54. Hu, J.; Zhao, W.; Wang, Y.; et al. The role of fluorine in polyanionic cathode materials for sodium-ion batteries. Small. Methods. 2025, e2402099.

55. Zhu, L.; Wang, H.; Sun, D.; Tang, Y.; Wang, H. A comprehensive review on the fabrication, modification and applications of Na3V2(PO4)2F3 cathodes. J. Mater. Chem. A. 2020, 8, 21387-407.

56. Bianchini, M.; Brisset, N.; Fauth, F.; et al. Na3V2(PO4)2F3 revisited: a high-resolution diffraction study. Chem. Mater. 2014, 26, 4238-47.

57. Bianchini, M.; Fauth, F.; Brisset, N.; et al. Comprehensive investigation of the Na3V2(PO4)2F3 - Na3V2(PO4)2F3 system by operando high resolution synchrotron X-ray diffraction. Chem. Mater. 2015, 27, 3009-20.

58. Chen, X.; Wu, Q.; Guo, P.; Liu, X. Rational design of two dimensional single crystalline Na3V2(PO4)2F3 nanosheets for boosting Na+ migration and mitigating grain pulverization. J. Chem. Eng. 2022, 439, 135533.

59. Deng, L.; Yu, F.; Xia, Y.; et al. Stabilizing fluorine to achieve high-voltage and ultra-stable Na3V2(PO4)2F3 cathode for sodium ion batteries. Nano. Energy. 2021, 82, 105659.

60. Gu, Z. Y.; Guo, J. Z.; Cao, J. M.; et al. An advanced high-entropy fluorophosphate cathode for sodium-ion batteries with increased working voltage and energy density. Adv. Mater. 2022, 34, e2110108.

61. Li, L.; Zhang, N.; Su, Y.; et al. Fluorine dissolution-induced capacity degradation for fluorophosphate-based cathode materials. ACS. Appl. Mater. Interfaces. 2021, 13, 23787-93.

62. Yang, Z.; Li, G.; Sun, J.; et al. High performance cathode material based on Na3V2(PO4)2F3 and Na3V2(PO4)2F3 for sodium-ion batteries. Energy. Storage. Mater. 2020, 25, 724-30.

63. Yi, H.; Ling, M.; Xu, W.; Li, X.; Zheng, Q.; Zhang, H. VSC-doping and VSU-doping of Na3V2-xTix(PO4)2F3 compounds for sodium ion battery cathodes: analysis of electrochemical performance and kinetic properties. Nano. Energy. 2018, 47, 340-52.

64. Deng, L.; Sun, G.; Goh, K.; et al. Facile one-step carbothermal reduction synthesis of Na3V2(PO4)2F3/C serving as cathode for sodium ion batteries. Electrochimica. Acta. 2019, 298, 459-67.

65. Zhang, B.; Dugas, R.; Rousse, G.; Rozier, P.; Abakumov, A. M.; Tarascon, J. M. Insertion compounds and composites made by ball milling for advanced sodium-ion batteries. Nat. Commun. 2016, 7, 10308.

66. Wang, M.; Huang, X.; Wang, H.; Zhou, T.; Xie, H.; Ren, Y. Synthesis and electrochemical performances of Na3V2(PO4)2F3/C composites as cathode materials for sodium ion batteries. RSC. Adv. 2019, 9, 30628-36.

67. Liu, Q.; Wang, D.; Yang, X.; et al. Carbon-coated Na3V2(PO4)2F3 nanoparticles embedded in a mesoporous carbon matrix as a potential cathode material for sodium-ion batteries with superior rate capability and long-term cycle life. J. Mater. Chem. A. 2015, 3, 21478-85.

68. Li, L.; Xu, Y.; Sun, X.; He, S.; Li, L. High capacity-favorable tap density cathode material based on three-dimensional carbonous framework supported Na3V2(PO4)2F3 nanoparticles. Chem. Eng. J. 2018, 331, 712-9.

69. Liu, Q.; Meng, X.; Wei, Z.; et al. Core/Double-shell structured Na3V2(PO4)2F3@C nanocomposite as the high power and long lifespan cathode for sodium-ion batteries. ACS. Appl. Mater. Interfaces. 2016, 8, 31709-15.

70. Jiang, T.; Chen, G.; Li, A.; Wang, C.; Wei, Y. Sol-gel preparation and electrochemical properties of Na3V2(PO4)2F3/C composite cathode material for lithium ion batteries. J. Alloys. Compd. 2009, 478, 604-7.

71. Eshraghi, N.; Caes, S.; Mahmoud, A.; Cloots, R.; Vertruyen, B.; Boschini, F. Sodium vanadium (III) fluorophosphate/carbon nanotubes composite (NVPF/CNT) prepared by spray-drying: good electrochemical performance thanks to well-dispersed CNT network within NVPF particles. Electrochimica. Acta. 2017, 228, 319-24.

72. Shen, C.; Long, H.; Wang, G.; Lu, W.; Shao, L.; Xie, K. Na3V2(PO4)2F3@C dispersed within carbon nanotube frameworks as a high tap density cathode for high-performance sodium-ion batteries. J. Mater. Chem. A. 2018, 6, 6007-14.

73. Leng, J.; Wang, Z.; Wang, J.; et al. Advances in nanostructures fabricated via spray pyrolysis and their applications in energy storage and conversion. Chem. Soc. Rev. 2019, 48, 3015-72.

74. Yi, H.; Lin, L.; Ling, M.; et al. Scalable and economic synthesis of high-performance Na3V2(PO4)2F3 by a solvothermal-ball-milling method. ACS. Energy. Lett. 2019, 4, 1565-71.

75. Hu, F.; Jiang, X. Superior performance of carbon modified Na3V2(PO4)2F3 cathode material for sodium-ion batteries. Inorg. Chem. Commun. 2021, 129, 108653.

76. Cai, Y.; Cao, X.; Luo, Z.; et al. Caging Na3V2(PO4)2F3 microcubes in cross-linked graphene enabling ultrafast sodium storage and long-term cycling. Adv. Sci. (Weinh). 2018, 5, 1800680.

77. Li, Y.; Liang, X.; Zhong, G.; et al. Fiber-shape Na3V2(PO4)2F3@N-Doped carbon as a cathode material with enhanced cycling stability for Na-ion batteries. ACS. Appl. Mater. Interfaces. 2020, 12, 25920-9.

78. Li, X.; Chen, W.; Qian, Q.; et al. Electrospinning-based strategies for battery materials. Adv. Energy. Mater. 2021, 11, 2000845.

79. Qiu, R.; Fei, R.; Guo, J.; et al. Encapsulation of Na3(VO)2(PO4)2F into carbon nanofiber as an superior cathode material for flexible sodium-ion capacitors with high-energy-density and low-self-discharge. J. Power. Sources. 2020, 466, 228249.

80. Shen, X.; Zhou, Q.; Han, M.; et al. Rapid mechanochemical synthesis of polyanionic cathode with improved electrochemical performance for Na-ion batteries. Nat. Commun. 2021, 12, 2848.

81. Minart, G.; Croguennec, L.; Weill, F.; Labrugère-Sarroste, C.; Olchowka, J. Increasing tap density of carbon-coated Na3V2(PO4)2F3 via mechanical grinding: good or bad idea? ACS. Appl. Energy. Mater. 2024, 7, 11334-42.

82. Serras, P.; Palomares, V.; Goñi, A.; Kubiak, P.; Rojo, T. Electrochemical performance of mixed valence Na3V2O2x(PO4)2F3-2x/C as cathode for sodium-ion batteries. J. Power. Sources. 2013, 241, 56-60.

83. Serras, P.; Palomares, V.; Goñi, A.; et al. High voltage cathode materials for Na-ion batteries of general formula Na3V2O2x(PO4)2F3-2x. J. Mater. Chem. 2012, 22, 22301.

84. Zhu, P.; Peng, W.; Guo, H.; et al. Toward high-performance sodium storage cathode: construction and purification of carbon-coated Na3V2(PO4)2F3 materials. J. Power. Sources. 2022, 546, 231986.

85. Jiang, N.; Zhang, L.; Cui, C.; Gao, L.; Yang, X. Synthesis and electrochemical performance of uniform carbon-coated Na3V2(PO4)2F3 using tannic acid as a chelating agent and carbon source. ACS. Appl. Energy. Mater. 2022, 5, 249-56.

86. Zhang, J.; Zhang, C.; Han, Y.; Zhao, X.; Liu, W.; Ding, Y. A surface-modified Na3V2(PO4)2F3 cathode with high rate capability and cycling stability for sodium ion batteries. RSC. Adv. 2024, 14, 13703-10.

87. Wang, S.; Li, J.; Xu, L.; et al. Manipulation of Na3V2(PO4)2F3 via aluminum doping to alter local electron states toward an advanced cathode for sodium-ion batteries. Rare. Met. 2024, 43, 4253-62.

88. Guan, J.; Zhou, S.; Zhou, J.; et al. Microwave-assisted hydrothermal synthesis of Na3V2(PO4)2F3 nanocuboid@reduced graphene oxide as an ultrahigh-rate and superlong-lifespan cathode for fast-charging sodium-ion batteries. ACS. Appl. Mater. Interfaces. 2024. DOI: 10.1021/acsami.4c01894.s001.

89. Al-Marri, A. H. Superior electrochemical properties of Na3V2(PO4)2F3/rGO composite cathode for high-performance sodium-ion batteries. J. Solid. State. Electrochem. 2024, 28, 2861-72.

90. Liang, K.; Zhao, H.; Li, J.; et al. Engineering crystal growth and surface modification of Na3V2(PO4)2F3 cathode for high-energy-density sodium-ion batteries. Small 2023, 19, e2207562.

91. Zhai, X.; Chen, X.; Zhang, Q.; et al. Temperature-dependent defect evolution and electrochemical performance enhancement of Na3V2(PO4)2F3. J. Alloys. Compd. 2023, 952, 170001.

92. Lin, Z. Phase formation in NaH2PO4-VOSO4-NaF-H2O system and rapid synthesis of Na3V2O2x(PO4)2F3-2x. Crystals 2024, 14, 43.

93. Wang, S.; Liang, K.; Li, J.; Huang, X.; Ren, Y. Surfactant-assisted synthesis of self-assembled Na3V2(PO4)2F3@C microspheres as the cathode for Na-ion batteries. Vacuum 2023, 211, 111894.

94. Moossa, B.; Abraham, J. J.; Ahmed, A. M.; Kahraman, R.; Al-qaradawi, S.; Shakoor, R. Synergistic effect of NASICON Na3V2(PO4)2F3 and 2D MXene for high-performance symmetric Sodium-ion batteries. Mater. Res. Bull. 2025, 182, 113173.

95. Guo, S.; Peng, J.; Sharma, N.; et al. Optimizing Sc-doped Na3V2(PO4)2F3/C as a high-performance cathode material for sodium-ion battery applications. Chem. Mater. 2025, 37, 1500-12.

96. Sun, C.; Zhang, L.; Xiong, X.; Deng, Z.; Sun, H.; Yang, X. Electronic/Ionic dual functional layer-coated Na3V2(PO4)2F3 cathode with high sodium storage performance. ACS. Sustainable. Chem. Eng. 2024, 12, 10892-904.

97. Yang, Y.; Xu, G. R.; Tang, A. P.; et al. Na3V2(PO4)2F3-decorated Na3V2(PO4)2F3 as a high-rate and cycle-stable cathode material for sodium ion batteries. RSC. Adv. 2024, 14, 11862-71.

98. Qin, Y.; Li, L.; Zhao, H.; et al. Effect of chelator content on the structural and electrochemical performance of Na3V2(PO4)2F3 by sol-gel preparation. CrystEngComm 2022, 24, 4519-26.

99. Mahato, S.; Das, S.; Gupta, D.; Biswas, K. Vanadium substituted Fe, Cr co-doped high performance C/Na3V2(PO4)2F3 cathode for sodium-ion batteries. J. Electroanal. Chem. 2024, 955, 118046.

100. Liang, K.; Zhao, H.; Li, J.; Huang, X.; Ren, Y. High-performance Na3V2(PO4)2F3 cathode obtained by a three-in-one strategy for self-sodium compensation, interface modification, and crosslinked carbon coatings. Appl. Surface. Sci. 2023, 615, 156412.

101. Hu, Z.; Zhang, R.; Fan, C.; et al. Synergistic effect, structural and morphology evolution, and doping mechanism of spherical Br-doped Na3V2(PO4)2F3/C toward enhanced sodium storage. Small 2022, 18, e2201719.

102. Zhou, Q.; Wang, Y.; Ou, R.; et al. Yolk-Shell Construction of Na3V2(PO4)2F3 with copper substitution microsphere as high-rate and long-cycling cathode materials for sodium-ion batteries. Small 2024, 20, e2310699.

103. Lei, L.; Sun, K.; Zhao, H.; Wang, C.; Wei, T. Large scale preparation of Na3V2(PO4)2F3 with cross-linked double carbon network for high energy density sodium ion batteries at -20 °C. J. Energy. Storage. 2024, 78, 109923.

104. Guo, R.; Li, W.; Lu, M.; et al. Na3V2(PO4)2F3@bagasse carbon as cathode material for lithium/sodium hybrid ion battery. Phys. Chem. Chem. Phys. 2022, 24, 5638-45.

105. Liang, M.; Li, W.; Yang, Y.; et al. Carbon Nanofiber/ Na3V2(PO4)2F3 particle composites as a self-standing cathode for high-voltage flexible sodium-ion batteries. ACS. Appl. Nano. Mater. 2023, 6, 22275-82.

106. Song, Z.; Liu, Y.; Guo, Z.; et al. Ultrafast synthesis of large-sized and conductive Na3V2(PO4)2F3 simultaneously approaches high tap density, rate and cycling capability. Adv. Funct. Mater. 2024, 34, 2313998.

107. Wang, J.; Jing, H.; Wang, X.; et al. Electrostatically shielded transportation enabling accelerated Na+ diffusivity in high-performance fluorophosphate cathode for sodium-ion batteries. Adv. Funct. Mater. 2024, 34, 2315318.

108. Ling, R.; Zhao, S.; Yang, C.; Qi, W. Three-dimensional ordered microporous Na3V2(PO4)2F3@C/carbon cloth as high-rate and stable flexible cathodes for Na-ion and Zn-ion batteries. Appl. Surf. Sci. 2023, 620, 156875.

109. Gu, Z. Y.; Guo, J. Z.; Sun, Z. H.; et al. Carbon-coating-increased working voltage and energy density towards an advanced Na3V2(PO4)2F3@C cathode in sodium-ion batteries. Sci. Bull. (Beijing). 2020, 65, 702-10.

110. Zhang, L. L.; Zhou, Y. X.; Li, T.; Ma, D.; Yang, X. L. Multi-heteroatom doped carbon coated Na3V2(PO4)2F3 derived from ionic liquids. Dalton. Trans. 2018, 47, 4259-66.

111. Li, F.; Zhao, Y.; Xia, L.; Yang, Z.; Wei, J.; Zhou, Z. Well-dispersed Na3V2(PO4)2F3@rGO with improved kinetics for high-power sodium-ion batteries. J. Mater. Chem. A. 2020, 8, 12391-7.

112. Yang, X.; Wang, X.; Zhen, W. Reversible Na+-extraction/insertion in nitrogen-doped graphene-encapsulated Na3V2(PO4)2F3@C electrode for advanced Na-ion battery. Ceram. Int. 2020, 46, 9170-5.

113. Hu, L.; Cheng, S.; Xiao, S.; et al. Dually decorated Na3V2(PO4)2F3 by carbon and 3D graphene as cathode material for sodium-ion batteries with high energy and power densities. ChemElectroChem 2020, 7, 3975-83.

114. Guo, H.; Hu, Y.; Zhang, X.; et al. Facile one-step hydrothermal synthesis of Na3V2(PO4)2F3/CNTs tetragonal micro-particles as high performance cathode material for Na-ion batteries. Front. Chem. 2019, 7, 689.

115. Kosova, N.; Rezepova, D. Mixed sodium-lithium vanadium fluorophosphates Na3-xLixV2(PO4)2F3: the origin of the excellent high-rate performance. J. Power. Sources. 2018, 408, 120-7.

116. Li, L.; Liu, X.; Tang, L.; Liu, H.; Wang, Y. Improved electrochemical performance of high voltage cathode Na3V2(PO4)2F3 for Na-ion batteries through potassium doping. J. Alloys. Compd. 2019, 790, 203-11.

117. Liu, W.; Yi, H.; Zheng, Q.; Li, X.; Zhang, H. Y-doped Na3V2(PO4)2F3 compounds for sodium ion battery cathodes: electrochemical performance and analysis of kinetic properties. J. Mater. Chem. A. 2017, 5, 10928-35.

118. Criado, A.; Lavela, P.; Pérez-vicente, C.; Ortiz, G.; Tirado, J. Effect of chromium doping on Na3V2(PO4)2F3@C as promising positive electrode for sodium-ion batteries. J. Electroanal. Chem. 2020, 856, 113694.

119. Guo, C.; Yang, J.; Cui, Z.; et al. In-situ structural evolution analysis of Zr-doped Na3V2(PO4)2F3 coated by N-doped carbon layer as high-performance cathode for sodium-ion batteries. J. Energy. Chem. 2022, 65, 514-23.

120. Cao, J.; Wang, Y.; Chen, Y.; et al. Improved sodium storage properties of co-doped Na3V2(PO4)2F3@graphene as anode material for sodium ion batteries. Ferroelectrics 2021, 584, 221-9.

121. Gu, Z. Y.; Guo, J. Z.; Sun, Z. H.; et al. Aliovalent-ion-induced lattice regulation based on charge balance theory: advanced fluorophosphate cathode for sodium-ion full batteries. Small 2021, 17, e2102010.

122. Zhang, Y.; Guo, S.; Xu, H. Synthesis of uniform hierarchical Na3V1.95Mn0.05(PO4)2F3@C hollow microspheres as a cathode material for sodium-ion batteries. J. Mater. Chem. A. 2018, 6, 4525-34.

123. Li, L.; Xu, Y.; Chang, R.; Wang, C.; He, S.; Ding, X. Unraveling the mechanism of optimal concentration for Fe substitution in Na3V2(PO4)2F3/C for sodium-ion batteries. Energy. Storage. Mater. 2021, 37, 325-35.

124. Gu, Z.; Guo, J.; Zhao, X.; et al. High-ionicity fluorophosphate lattice via aliovalent substitution as advanced cathode materials in sodium-ion batteries. InfoMat 2021, 3, 694-704.

125. Missaoui, K.; Ferchichi, K.; Amdouni, N.; et al. Polyaniline-coated Na3V2(PO4)2F3 cathode enables fast sodium ion diffusion and structural stability in rechargeable batteries. ACS. Appl. Mater. Interfaces. 2024, 16, 50550-60.

126. Mani, K. P. L.; Venkatesh, M.; Nandy, S.; et al. Outstanding specific energy achieved via reversible cycling of V4+/V2+ redox couple in N-doped carbon coated Na3V2(PO4)2F3: an ex-situ XRD, XPS and XAS study. Mater. Today. Energy. 2025, 48, 101802.

127. Sun, C.; Zhang, L.; Deng, Z.; Yan, B.; Gao, L.; Yang, X. PTFE-derived carbon-coated Na3V2(PO4)2F3 cathode material for high-performance sodium ion battery. Electrochimica. Acta. 2022, 432, 141187.

128. Hu, Y.; Chen, P.; Liu, F.; et al. Dual-anion ether electrolyte enables stable high-voltage Na3V2(PO4)2F3 cathode under wide temperatures. J. Power. Sources. 2024, 602, 234405.

129. Zhao, W.; Wang, W.; Hu, G.; et al. Facial construction of high rate Na3V2(PO4)2F3/C microspheres with fluorocarbon layer by deep-eutectic solvent synthesis. Electrochimica. Acta. 2023, 440, 141718.

130. Yang, X.; Wang, M.; Xiang, X.; Liu, S.; Chen, C. An open-system synthesis approach to achieve high-rate Na3(VO)2(PO4)2F/C microcubes cathode for sodium-ion batteries. J. Electroanal. Chem. 2024, 956, 118088.

131. Zhang, Y.; Song, W.; Tang, Y.; Jia, D.; Huang, Y. Amylopectin-assisted fabrication of in situ carbon-coated Na3V2(PO4)2F3 nanosheets for ultra-fast sodium storage. ACS. Appl. Mater. Interfaces. 2022, 14, 40812-21.

132. Wang, M.; Wang, Y.; Xin, Y.; Liu, Q.; Wu, F.; Gao, H. Nitrogen-doped carbon coated Na3V2(PO4)2F3 derived from polyvinylpyrrolidone as a high-performance cathode for sodium-ion batteries. ACS. Appl. Energy. Mater. 2023, 6, 4453-61.

133. Sun, C.; Zhang, L. L.; Deng, Z. R.; Sun, H. B.; Yang, X. L. Achieving high-performance Na3V2(PO4)2F3 cathode material through a bifunctional N-doped carbon network. ACS. Appl. Mater. Interfaces. 2024, 16, 35179-89.

134. Xu, J.; Tang, A.; Wen, Q.; et al. N/S dual-doped KB-decorated Na3V2(PO4)2F3 as high-performance cathode for advanced sodium storage properties. Ionics 2024, 30, 7037-49.

135. Wang, A.; Tian, Z.; Li, X.; Chai, Y.; Wang, N. Codoping of carbon and boron composition in Na3V2(PO4)2F3 affects its sodium storage properties. J. Electroanal. Chem. 2024, 974, 118741.

136. Zhang, X.; Tian, H.; Zhang, Y.; Cai, Y.; Yao, X.; Su, Z. Diatomic-doped carbon layer decorated Na3V2(PO4)2F3 as a durable ultrahigh-stability cathode for sodium ion batteries. New. J. Chem. 2023, 47, 9611-7.

137. Yu, X.; Lu, T.; Li, X.; et al. Realizing outstanding electrochemical performance with Na3V2(PO4)2F3 modified with an ionic liquid for sodium-ion batteries. RSC. Adv. 2022, 12, 14007-17.

138. Yu, X.; Lu, T.; Li, X.; et al. Ionic liquid-acrylic acid copolymer derived nitrogen-boron codoped carbon-covered Na3V2(PO4)2F3 as cathode material of high-performance sodium-ion batteries. Langmuir 2022, 38, 7815-24.

139. Zhang, X.; Tian, H.; Cai, Y.; Wang, L.; Yao, X.; Su, Z. Effects of nitrogen and sulfur atom regulation on electrochemical properties of Na3V2(PO4)2F3 cathode material for Na-ion batteries. Ceram. Int. 2022, 48, 36129-35.

140. Tang, K.; Tian, H.; Zhang, Y.; et al. Multilevel carbon composite construction of NASICON-type NaVPO4F/C/CNT cathode material for enhanced-performance sodium-ion batteries. J. Mater. Chem. C. 2025, 13, 6605-13.

141. Ma, W. L.; Zhou, Y.; Zhao, X. W.; et al. Ultra-fast-charging, long-duration, and wide-temperature-range sodium storage enabled by multiwalled carbon nanotube-hybridized biphasic polyanion-type phosphate cathode materials. ACS. Appl. Mater. Interfaces. 2024, 16, 34819-29.

142. Li, L.; Qin, Y.; Zhang, S.; et al. Ion transport through carbon nanotubes enable highly crystalline Na3V2(PO4)2F3 cathode for ultra-stable sodium-ion storage. J. Power. Sources. 2023, 576, 233226.

143. Zhang, Q.; Sun, X.; Liu, K.; Xu, Q.; Zheng, S.; Dai, S. Synergistic coupling effect of electronic conductivity and interphase compatibility on high-voltage Na3V2(PO4)2F3 cathodes. ACS. Sustainable. Chem. Eng. 2023, 11, 12992-3001.

144. Gao, J.; Tian, Y.; Ni, L.; et al. Robust cross-linked Na3V2(PO4)2F3 full sodium-ion batteries. Energy. Environ. Mater. 2024, 7, e12485.

145. Qin, M.; Qin, N.; Lei, M.; et al. Construction of Na3V2(PO4)2F3@C/CNTs nanocomposites with three-dimensional conductive network as cathode materials for sodium-ion batteries. J. Electroanal. Chem. 2022, 920, 116613.

146. He, J.; Tao, T.; Yang, F.; Sun, Z.; Huang, H. Phase-manipulated hierarchically core-shell Na3(VO1-xPO4)2F1+2x(0 ≤ x ≤ 1)@Na3V2(PO4)3 and its synergistic effect with conformally wrapped reduced graphene oxide framework towards high-performance cathode for sodium-ion batteries. Mater. Today. Phys. 2022, 27, 100813.

147. Xu, S.; Zhu, Y.; Li, X.; et al. PVA-regulated construction of 3D rGO-hosted Na3V2(PO4)2F3 for fast and stable sodium storage. J. Energy. Chem. 2024, 99, 100-9.

148. Huang, Q.; Shao, L.; Shi, X.; et al. Na3V2O2(PO4)2F nanoparticles@reduced graphene oxide: a high-voltage polyanionic cathode with enhanced reaction kinetics for aqueous zinc-ion batteries. Chem. Eng. J. 2023, 468, 143738.

149. Ou, J.; Wang, H.; Deng, H.; Li, B.; Zhang, H. Hydrothermally prepared composite of Na3V2(PO4)2F3 with gelatin and graphene used as a high-performance sodium ion battery cathode. J. Alloys. Compd. 2022, 926, 166857.

150. Shi, C.; Xu, J.; Tao, T.; et al. Zero-Strain Na3V2(PO4)2F3@Rgo/CNT composite as a wide-temperature-tolerance cathode for Na-ion batteries with ultrahigh-rate performance. Small. Methods. 2024, 8, e2301277.

151. Zheng, Q.; Ni, X.; Lin, L.; et al. Towards enhanced sodium storage by investigation of the Li ion doping and rearrangement mechanism in Na3V2(PO4)2F3 for sodium ion batteries. J. Mater. Chem. A. 2018, 6, 4209-18.

152. Kuang, Q.; Zhao, Y.; Liang, Z. Synthesis and electrochemical properties of Na-doped Li3V2(PO4)3 cathode materials for Li-ion batteries. J. Power. Sources. 2011, 196, 10169-75.

153. Wu, Q.; Ma, Y.; Zhang, S.; et al. Achieving a rapid Na+ migration and highly reversible phase transition of NASICON for sodium-ion batteries with suppressed voltage hysteresis and ultralong lifespan. Small 2024, 20, e2404660.

154. Fu, W.; Li, B.; Wang, P.; Lin, Z.; Zhu, K. A high-entropy carbon-coated Na3V1.9(Mg, Cr, Al, Mo, Nb)0.1(PO4)2F3 cathode for superior performance sodium-ion batteries. Ceram. Int. 2024, 50, 16166-71.

155. Yang, J.; Liu, N.; Jiang, G.; et al. Synthesis and investigation of sodium storage properties in Na3V1.9Fe0.1(PO4)2F3@N-CNTs cathode material for sodium ion batteries. Chem. Eng. J. 2024, 485, 149834.

156. Puspitasari, D. A.; Patra, J.; Hung, I.; Bresser, D.; Lee, T.; Chang, J. Optimizing the Mg doping concentration of Na3V2-xMgx(PO4)2F3/C for enhanced sodiation/desodiation properties. ACS. Sustainable. Chem. Eng. 2021, 9, 6962-71.

157. Ren, K.; Qiu, J.; Liu, H.; Song, H.; Li, Q.; Li, J. Low-valence Mg2+ doping suppresses irreversible phase transition of sodium-rich fluorophosphate upon additional Na+ deintercalation. ACS. Appl. Energy. Mater. 2025, 8, 3066-73.

158. Olchowka, J.; Nguyen, L. H. B.; Broux, T.; et al. Aluminum substitution for vanadium in the Na3V2(PO4)2F3 and Na3V2(PO4)2FO2 type materials. Chem. Commun. (Camb). 2019, 55, 11719-22.

159. Pineda-Aguilar, N.; Gallegos-Sánchez, V. J.; Sánchez, E. M.; Torres-gonzález, L. C.; Garza-tovar, L. L. Aluminum doped Na3V2(PO4)2F3 via sol-gel Pechini method as a cathode material for lithium ion batteries. J. Sol-Gel. Sci. Technol. 2017, 83, 405-12.

160. Wang, J.; Liu, Q.; Cao, S.; Zhu, H.; Wang, Y. Boosting sodium-ion battery performance with binary metal-doped Na3V2(PO4)2F3 cathodes. J. Colloid. Interface. Sci. 2024, 665, 1043-53.

161. Chen, Q.; Gong, F.; Pan, S.; Chen, W. Effects of Bi doping on the electrochemical performance of Na3V2(PO4)3F3 cathode material for sodium ion batteries. Solid. State. Ionics. 2024, 414, 116621.

162. Puspitasari, D. A.; Patra, J.; Hernandha, R. F. H.; et al. Enhanced electrochemical performance of Ca-doped Na3V2(PO4)2F3/C cathode materials for sodium-ion batteries. ACS. Appl. Mater. Interfaces. 2024, 16, 496-506.

163. Silva, C. H. P.; de, O. L. J. L.; Rezende, M. V. D. S. Intrinsic defect and deformation on local structure caused by rare-earth ions in the Na3V2(PO4)2F3 cathode material. Physica. Status. Solidi. (b). 2024, 261, 2300348.

164. Ghosh, S.; Barman, N.; Mazumder, M.; Pati, S. K.; Rousse, G.; Senguttuvan, P. High capacity and high-rate NASICON-Na3.75V1.25Mn0.75(PO4)3 cathode for Na-ion batteries via modulating electronic and crystal structures. Adv. Energy. Mater. 2020, 10, 1902918.

165. Li, T.; Yang, D.; Xu, G.; et al. A novel NASICON-typed Na3V1.96Cr0.03Mn0.01(PO4)2F3 cathode for high-performance Na-ion batteries. J. Energy. Storage. 2024, 104, 114596.

166. Cai, C.; Liu, Q.; Hu, Z.; et al. Construction of superior performance Na3V2-xCrx(PO4)2F3/C cathode by homovalent doping strategy toward enhanced sodium ion storage. J. Power. Sources. 2023, 571, 233080.

167. Yi, X.; Luo, H.; Zhou, Y.; et al. Effect of Cr3+ doping on the electrochemical performance of Na3V2(PO4)2F3/C cathode materials for sodium ion battery. Electrochimica. Acta. 2023, 437, 141491.

168. Gu, Z.; Heng, Y.; Guo, J.; et al. Nano self-assembly of fluorophosphate cathode induced by surface energy evolution towards high-rate and stable sodium-ion batteries. Nano. Res. 2023, 16, 439-48.

169. Su, R.; Zhu, W.; Liang, K.; et al. Mnx+ substitution to improve Na3V2(PO4)2F3-based electrodes for sodium-ion battery cathode. Molecules 2023, 28, 1409.

170. Tong, S.; Pan, H.; Liu, H.; et al. Titanium doping induced the suppression of irreversible phase transformation at high voltage for V-based phosphate cathodes of Na-ion batteries. ChemSusChem 2023, 16, e202300244.

171. Nongkynrih, J.; Sengupta, A.; Modak, B.; Mitra, S.; Tyagi, A.; Dutta, D. P. Enhanced electrochemical properties of W-doped Na3V2(PO4)2F3@C as cathode material in sodium ion batteries. Electrochimica. Acta. 2022, 415, 140256.

172. Zhu, L.; Zhang, Q.; Sun, D.; et al. Engineering the crystal orientation of Na3V2(PO4)2F3@rGO microcuboids for advanced sodium-ion batteries. Mater. Chem. Front. 2020, 4, 2932-42.

173. Park, S.; Song, J.; Kim, S.; et al. Phase-pure Na3V2(PO4)2F3 embedded in carbon matrix through a facile polyol synthesis as a potential cathode for high performance sodium-ion batteries. Nano. Res. 2019, 12, 911-7.

174. Liu, S.; Cao, X.; Zhang, Y.; et al. Carbon quantum dot modified Na3V2(PO4)2F3 as a high-performance cathode material for sodium-ion batteries. J. Mater. Chem. A. 2020, 8, 18872-9.

175. Ponrouch, A.; Dedryvère, R.; Monti, D.; et al. Towards high energy density sodium ion batteries through electrolyte optimization. Energy. Environ. Sci. 2013, 6, 2361.

176. Sadan, M. K.; Kim, H.; Kim, C.; et al. Enhanced rate and cyclability of a porous Na3V2(PO4)3 cathode using dimethyl ether as the electrolyte for application in sodium-ion batteries. J. Mater. Chem. A. 2020, 8, 9843-9.

177. Hwang, J.; Matsumoto, K.; Hagiwara, R. Electrolytes toward high-voltage Na3V2(PO4)2F3 positive electrode durable against temperature variation. Adv. Energy. Mater. 2020, 10, 2001880.

178. Liu, F.; Zong, J.; Liang, Y.; et al. Ordered vacancies as sodium ion micropumps in Cu-deficient copper indium diselenide to enhance sodium storage. Adv. Mater. 2024, 36, e2403131.

179. Li, J.; Liang, Z.; Jin, Y.; et al. A high-voltage cathode material with ultralong cycle performance for sodium-ion batteries. Small. Methods. 2024, 8, e2301742.

180. Xie, K.; Ji, Y.; Yang, L.; Pan, F. Electrolyte design strategies to construct stable cathode-electrolyte interphases for high-voltage sodium-ion batteries. Adv. Energy. Mater. 2025, 15, 2405301.

181. Alptekin, H.; Au, H.; Olsson, E.; et al. Elucidation of the solid electrolyte interphase formation mechanism in micro-mesoporous hard-carbon anodes. Adv. Mater. Inter. 2022, 9, 2101267.

182. Zhang, J.; Li, J.; Jia, G.; Wang, H.; Wang, M. Improving Na3V2(PO4)2F3 half-cell performance with NaBF4-enhanced sodium difluoro(oxalato)borate electrolyte. J. Energy. Chem. 2025, 102, 340-52.

183. Liang, H.; Liu, H.; Guo, J.; et al. Self-purification and silicon-rich interphase achieves high-temperature (70 °C) sodium-ion batteries with nonflammable electrolyte. Energy. Storage. Mater. 2024, 66, 103230.

184. Liang, H. J.; Liu, H. H.; Zhao, X. X.; et al. Electrolyte chemistry toward ultrawide-temperature (-25 to 75 °C) sodium-ion batteries achieved by phosphorus/silicon-synergistic interphase manipulation. J. Am. Chem. Soc. 2024, 146, 7295-304.

185. Ren, H.; Zhang, X.; Liu, Q.; Tang, W.; Liang, J.; Wu, W. Fully-printed flexible aqueous rechargeable sodium-ion batteries. Small 2024, 20, e2312207.

186. Desai, P.; Forero-Saboya, J.; Meunier, V.; et al. Mastering the synergy between Na3V2(PO4)2F3 electrode and electrolyte: a must for Na-ion cells. Energy. Storage. Mater. 2023, 57, 102-17.

187. Zheng, Y.; Sun, M.; Yu, F.; et al. Utilizing weakly-solvated diglyme-based electrolyte to achieve a 10,000-cycles durable Na3V2(PO4)2F3 cathode endured at -20 °C. Nano. Energy. 2022, 102, 107693.

188. Wang, X.; Yang, C.; Yao, L.; Wang, Y.; Jiang, N.; Liu, Y. Anion/Cation solvation engineering for a ternary low-concentration electrolyte toward high-voltage and long-life sodium-ion batteries. Adv. Funct. Mater. 2024, 34, 2315007.

189. Jiang, M.; Li, T.; Qiu, Y.; et al. Electrolyte design with dual -C≡N groups containing additives to enable high-voltage Na3V2(PO4)2F3-based sodium-ion batteries. J. Am. Chem. Soc. 2024, 146, 12519-29.

190. Hwang, J.; Aoyagi, I.; Takiyama, M.; Matsumoto, K.; Hagiwara, R. Inhibition of aluminum corrosion with the addition of the tris(pentafluoroethyl)trifluorophosphate anion to a sulfonylamide-based ionic liquid for sodium-ion batteries. J. Electrochem. Soc. 2022, 169, 080522.

191. Yan, Y.; Xu, J.; Cao, J.; et al. Fluorinated ionic liquid mediated nanostructure with enhanced conductivity and fluorine retention in Na3V2(PO4)2F3 cathode toward high-performance sodium-ion batteries. Chem. Eng. J. 2024, 498, 155640.

192. Li, Z.; Qiu, L.; Li, P.; et al. Exposing the (002) active facet by reducing surface energy for a high-performance Na3V2(PO4)2F3 cathode. J. Mater. Chem. A. 2024, 12, 7777-87.

193. Zhang, S.; Wang, J.; Chen, K.; et al. Aromatic ketones as mild presodiating reagents toward cathodes for high-performance sodium-ion batteries. Angew. Chem. Int. Ed. Engl. 2024, 63, e202317439.

194. Yun, D. H.; Song, J.; Kim, J.; et al. A binder-driven cathode-electrolyte interphase via a displacement reaction for high voltage Na3V2(PO4)2F3 cathodes in sodium-ion batteries. J. Mater. Chem. A. 2023, 11, 5540-7.

195. Zhang, Z.; Zhang, R.; Rajagopalan, R.; et al. A high-capacity self-sacrificial additive based on electroactive sodiated carbonyl groups for sodium-ion batteries. Chem. Commun. (Camb). 2022, 58, 8702-5.

196. Forero-saboya, J.; Desai, P.; Healy, C. R.; et al. Influence of formation temperature on cycling stability of sodium-ion cells: a case study of Na3V2(PO4)2F3|HC Cells. J. Electrochem. Soc. 2023, 170, 100529.

197. Komayko, A. I.; Shraer, S. D.; Fedotov, S. S.; Nikitina, V. A. Advantages of a solid solution over biphasic intercalation for vanadium-based polyanion cathodes in Na-ion batteries. ACS. Appl. Mater. Interfaces. 2023, 15, 43767-77.

198. Goloviznina, K.; Bendadesse, E.; Sel, O.; Tarascon, J. M.; Salanne, M. Disclosing the interfacial electrolyte structure of Na-insertion electrode materials: origins of the desolvation phenomenon. ACS. Appl. Mater. Interfaces. 2023, 15, 59380-8.

199. Semykina, D. O.; Sharafutdinov, M. R.; Kosova, N. V. Understanding of the mechanism and kinetics of the fast solid-state reaction between NaF and VPO4 to form Na3V2(PO4)2F3. Inorg. Chem. 2022, 61, 10023-35.

200. Deng, L.; Yu, F. D.; Sun, G.; et al. Constructing stable anion-tuned electrode/electrolyte interphase on high-voltage Na3V2(PO4)2F3 cathode for thermally-modulated fast-charging batteries. Angew. Chem. Int. Ed. Engl. 2022, 61, e202213416.

201. Zhang, Y.; Xun, J.; Zhang, K.; Zhang, B.; Xu, H. 2D-lamellar stacked Na3V2(PO4)2F3@RuO2 as a high-voltage, high-rate capability and long-term cycling cathode material for sodium ion batteries. J. Mater. Chem. A. 2022, 10, 11163-71.

202. S., Biswas, K. Boosting sodium-ion battery performance with vanadium substituted Fe, Ni dual doped fluorophosphate cathode over a wide temperature range. J. Power. Sources. 2025, 626, 235734.

Cite This Article

Review
Open Access
Recent research progress in the application of Na3V2(PO4)2F3 cathode materials in sodium-ion batteries: synthesis, modification, and battery optimization
Yining Liang, ... Zhipeng SunZhipeng Sun

How to Cite

Download Citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click on download.

Export Citation File:

Type of Import

Tips on Downloading Citation

This feature enables you to download the bibliographic information (also called citation data, header data, or metadata) for the articles on our site.

Citation Manager File Format

Use the radio buttons to choose how to format the bibliographic data you're harvesting. Several citation manager formats are available, including EndNote and BibTex.

Type of Import

If you have citation management software installed on your computer your Web browser should be able to import metadata directly into your reference database.

Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.

Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.

About This Article

© The Author(s) 2025. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
67
Downloads
2
Citations
0
Comments
0
0

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at [email protected].

0
Download PDF
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Energy Materials
ISSN 2770-5900 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/