Low-Dimensional Vanadium-Based High-Voltage Cathode Materials: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Contributor:

Owing to their rich structural chemistry and unique electrochemical properties, vanadium-based materials, especially the low-dimensional ones, are showing promising applications in energy storage and conversion.

  • vanadium-based materials
  • low-dimensional
  • nanomaterials
  • high-voltage cathodes
  • alkali-ion batteries
  • alkali-metal-ion batteries

1. Introduction

Vanadium-based materials have been considered one of the most promising cathode candidates for next-generation secondary batteries, especially sodium-ion batteries (SIBs) and potassium-ion batteries (PIBs), due to the merits of rich structural chemistry, high voltage output (up to over 4.0 V), cost-effectiveness, and sustainability [1][2][3]. The multivalent state of vanadium and the open framework or layered structure of vanadium-based materials are significantly beneficial to their superior performances, not only on the aspects of (rate) capacities but also the ultralong cycling stability of high-voltage alkali-ion batteries (AIBs, A = Li, Na, and K; namely, LIBs, SIBs, and PIBs) [3][4]. Polyanionic materials, e.g., phosphates, fluorophosphates, and pyrophosphates, mixed phosphates, are structurally diverse compounds, and the vanadium-containing phosphates and fluorophosphates such as Na3V2(PO4)3 and Na3V2(PO4)2F3 are among the most intensively studied high-voltage cathode candidate materials for promising commercial applications [5][6][7]. Low-dimensional (e.g., 0D, 1D, and 2D) materials possess unique advantages, including a high surface-area-to-volume ratio, shortened ion diffusion lengths, a larger electrolyte–electrode contact area, reduced charge–discharge time, enhanced buffering against stress and volume change, and efficient electron transport along the longitudinal direction (especially for 1D materials) [8][9].

2. Compositions, Structures, and Methods

2.1. Chemical Composition and Crystal Structure

Vanadium-based materials are generally layered or tunnel-type materials and thus are typically intercalation hosts for alkali ions, characterized by their intercalation-deintercalation (or insertion-extraction) mechanism and electrochemically long plateaus as cathode materials for alkali-ion batteries [10][11]. In operando X-ray diffraction (also called in situ XRD) is an effective way to analyze the underlying crystal structure during cell cycling [12][13], thus facilely relating the electrochemical processes with crystal structure and phase transformation for revealing the (de)intercalation mechanisms, degradation mechanisms, and optimizing the performances [14][15].
Vanadium oxides, vanadates, and vanadium phosphates are among the high-voltage cathode materials and have received intensive attention for electrochemical energy storage and conversion, especially for secondary batteries, in the last two decades [3]. And these vanadium-based materials are evaluated in detail, as follows:
Oxygen-free vanadium-based chalcogenides, nitrides, and carbides (e.g., MXenes) are novel 2D materials with high-capacity, high-conductivity, and/or high-stability advantages; however, they are usually comparatively inferior in voltage output or generally investigated as electrode materials for supercapacitors [16][17][18]. Polyanion-based phosphates/fluorophosphates [19][20] and layered transition-metal compounds (e.g., different types of O2, O3, P2, and P3, referring to edge- and face-sharing structures [21][22][23][24]) are characterized by their sodium super ionic conductor (NASICON)-type structures for promising high-voltage cathode materials. And for vanadium-based phosphates/fluorophosphates, (transition-)metal doping/substitution (e.g., Mn, Fe, Ni, Cr, Ce, Ti, Al, K) [6][25][26][27][28] and multivalent anion substitution (e.g., partially replacing PO43− with SiO44−) [29] are also emerging recently for higher performances (e.g., lower cost, higher capacities, and improved rate performance with enhanced ion/electron transport) [30].
To be specific, vanadium bronzes (or named vanadium oxide bronzes, abbr. VOBs, MxVyOz, or MxV2O5, M = Li, Na, K, Mg, Ca, Zn, NH4, Ag, etc.) are typical layered materials with triclinic vanadium–oxygen coordination polyhedral crystal structure and can be identified as vanadium oxides with accommodated electron-donating cations [31][32]. VOBs have raised increasing interest as cathode materials due to their intrinsic large interlayer spacing, mixed valences, high capacity, enhanced electronic conductivity, and structural stability for facilitated intercalation not only in lithium-based batteries but also in non-lithium-based batteries [31][32]. Layered sodium vanadium oxide, e.g., NaV6O15 with nanoflakes self-assembled microflower structure, can exhibit a plateau of ~2.75 V in the potential window of 4.0–1.5 V, and it can deliver a high capacity of 126 mA h g−1 at 100 mA g−1 as a cathode material for SIBs, as well as a capacity retention of 87% after 2000 cycles at a high current density of 5 A g−1 [33]. While for monoclinic NaVO3, the Na+ (de)intercalation does not affect the crystal structure and little change occurs for the a and b lattice parameters (0.13% and 0.19%, respectively), which corresponds to a significantly high specific capacity of 245 mA h g−1 in the potential range of 4.7–1.2 V (vs. Na+/Na), contributed by the cationic V5+/4+ and anionic O/O2− redox couples (during discharge process, and the latter redox reaction is partially reversible) [15]; although the possible operating voltage could be higher using density functional theory (DFT) prediction based on the dominant oxygen redox reaction [34].
The vanadium-containing phosphates can exchange even more than one electron per transition metal due to the wide range of oxidation states (+2 to +5) and the diverse polyhedra formed thereof (e.g., tetrahedra, octahedra, pyramids). Two typical redox couples of V3+/V4+ and V4+/V5+ (sometimes involving V3+/V2+ at lower voltages, e.g., below 2 V) have been frequently reported in these various vanadium phosphates. Of which, classic Li3V2(PO4)3 and Li-rich layered vanadium mixed-phosphates such as Li9V3(P2O7)3(PO4)2 and their derivatives (e.g., Na, Mg-substituted/doped) have been intensively investigated for their high-voltage plateaus (~4.5 V) and high capacities as cathodes of LIBs [35]. NASICON-type Na3V2(PO4)3 and its fluorophosphates (e.g., Na3V2(PO4)2F3) are also intensively studied for SIB cathodes, which both show long-term cycling stability, rate, and high-voltage performances [12].
Fluorophosphates (substituted PO43− with high-electronegativity F) are a novel family of promising cathode materials with even higher voltage output compared to phosphates [7]. Vanadium fluorophosphates with long and high-voltage plateaus and high theoretical capacities have been intensively investigated for LIBs, SIBs, and PIBs, e.g., NaVPO4F (theoretical capacity 143 mA h g−1) [36], Na3V2O2(PO4)2F (space group of I4/mmm, lattice parameters of e.g., a = b = 6.4958 Å, c = 10.61366 Å) [19]. These tetrahedral [PO4] and octahedral [VO5F]/[VO4F2] units were interconnected by an O atom in the ab-plane (i.e., (002) plane), which can thus be considered as pseudolayered structures with intercalated Na ions on the ab-planes [19][36]. The growth of these crystals by the hydrothermal method can be summarized as follows: under appropriate reaction conductions, the positive Na+ ions are immediately adsorbed onto the negative molecular clusters, followed by self-assembling along the normal direction of the basal ab-plane with the Na ions as pillars to stabilize the structure. The driving force is attributed to the much higher interfacial free energy of the ab-plane perpendicular to the c-axis than that of the other surfaces; thus, an anisotropic crystal structure such as 1D nanowires or nanorods usually forms in high-chemical-potential surroundings [19]. The possible Na+ diffusion paths are mainly between the ab-planes from a Na1 site to an adjacent Na1 site, since the activation barriers along the c-axis direction are much higher than those in the ab-plane (2.248 vs. 0.415 eV, calculated by first principles) [19]. And for 1D-structured materials, the preferred growth along the c-axis direction is favorable to the high-rate diffusion of Na+ ions since the diffusion paths are perpendicular to the nanowire/nanorod growth direction [19].

2.2. Structure, Synthesis, and Electrode Design

As it is known, particle size and intrinsic electrical conductivity affect the diffusion length and electronic/ionic transport; thus, crystal growth and assembly are of great importance. Well-designed nanostructured materials have been considered effective strategies for enhanced intercalation/deintercalation rates due to the shortened ion transport distance and more exposed crystal facets (high surface-to-bulk ratio). A lot of attention has been devoted to the synthesis of nanoscale low-dimensional cathode materials, including 0D nanoparticles [37], 1D nanorods/nanowires [9][38][39], and 2D nanoplates/nanosheets [40], to enhance the rate capability and cycle performance. Nanoparticles herein are classified into 0D materials, and micrometer-sized particles are not intentionally included in this research focusing on low-dimensional materials. To enhance the kinetics of larger alkali-ion (e.g., Na+, K+) transfer in the cathode, strategies including nanostructuring (decreasing the crystalline size), tuning structures/morphologies, and doping are typically adopted. The synthesis methods mainly include electrospinning (1D, 3D) [41], ball-milling, solid-state reaction (calcination/annealing) or carbothermal reduction (CTR), sol–gel [42], hydrothermal [37][38], etc. Of which, ball-milling and stoichiometric solid-state reactions are intensively used; however, the one-step high-energy ball milling and sol-gel method, together with annealing, i.e., synergistic methods, are promising for future low-cost approaches with high performances. Solid-state reaction and hydrothermal method are typical strategies to realize high-voltage cathode materials with high performance, e.g., solution route followed by solid-state reaction or ball milling followed by solid-state reaction. The high-energy ball milling (i.e., rapid mechanochemical synthesis by a 3D ball-milling machine [43]) may not need further solid-state reaction. Hydrothermal/solvothermal methods will render samples with higher crystallinity and usually no further annealing/calcination is needed.
Beyond the self-assembly featured in sol–gel or hydrothermal methods, some synthesis methods also determine the materials and electrode structures; e.g., electrospinning can produce 1D materials and the 3D aggregates thereof for free-standing/flexible additive-free cathodes of advanced batteries [36][44]. Electrospinning is a facile way to fabricate 1D nanostructures and the crosslinked 3D mat-like framework and is an ideal candidate for cathode materials with sufficient conductivity both for electrons and ions; however, the mass loading, active material ratio, and strength are also of great concern and may not be met for practical applications. And for electrospun 1D materials, active materials inside are in fact not 1D materials but exist in the form of (isolated) nanoparticles dispersed in the conductive backbones (usually carbon-based). Furthermore, some V-based nanoparticles may be in situ coated by carbon layers and/or embedded into 3D conducting networks (e.g., porous graphene network, carbon nanotube framework), which may have superior rate performances up to 100C or even 500C as well as high cycling stability up to even 10,000 cycles [42][45][46].

2.3. Electrochemical Evaluation

For higher voltage plateaus, a lower vanadium valence (e.g., 3+, 4+) with a specific crystal lattice is usually a critical requirement. Tunnel-structured vanadium phosphates, vanadium fluorophosphates, and layer-structured vanadium oxide bronzes are typical vanadium-containing cathode materials that are showing great promise for ultimately practical application.
Vanadium phosphates and fluorophosphates are intensively investigated as high-voltage (together with flat potential plateaus and very low potential hysteresis) cathode materials for alkali-ion batteries [1][45]. These phosphates/fluorophosphates are showing typical flat potential plateaus. Most of them are generally not only showing high-potential plateaus in 5.0–3.0 V but also low-potential plateaus in 2.0–1.0 V (i.e., big voltage difference, due to V3+/V4+/V5+ redox couple and V3+/V2+(/V1+) redox couple, respectively). Thus they are amphoteric and could be working as either cathodes or anodes, such as Li3V2(PO4)3, Na3V2(PO4)3 [37], LiVPO4F, β-LiVP2O7 [47], β-NaVP2O7 [48], and Li9V3(P2O7)3(PO4)2 [49]; symmetric batteries can therefore be configured by using the same material for cathode and anode simultaneously.
For typical polycationic compounds such as Na3V2(PO4)3, a very flat potential plateau (~3.4 V vs. Na+/Na) is seen, corresponding to the redox reaction between V3+/V4+ with two Na+ extractions/insertions, i.e., a typical two-phase reaction of Na3V2(PO4)3 ↔ NaV2(PO4)3, showing an ultralow potential hysteresis of 5 mV [45][50]. These V-based phosphates are usually isostructural to their Fe-based counterparts, and these Na-based phosphates are also isostructural to their Li-based counterparts, which could be prepared by facile electrochemical exchange of Na+ for Li+ [47]. Although a higher charging voltage will render them with a temporarily elevated voltage/capacity along with a three-phase transition and unexpected solid solution behavior (e.g., a three Li-ions extraction compared to only two Li-ions extraction in the normal two-phase transition for monoclinic α-Li3V2(PO4)3) [13], it is not stable and unsustainable, i.e., the cycling stability decreases with an increase in upper cut-off voltages due to the irreversible unit cell volume expansion, increased amount of V5+ and destruction of carbon layer coating on the surface of active materials [51]. Sodium vanadium pyrophosphates such as Na7V3(P2O7)4 and Na2VOP2O7 [52] possess a higher redox potential as the vanadium-based cathode of SIBs, an 4V-class electrode. Taking Na7V3(P2O7)4 as an example, the unique structure with octahedral VO6 and connected P2O7 groups in quasi-layers, as well as the increased inductive effect thereof, contributes to the higher redox potential and output voltage. The electrochemical redox mechanism is also based on the vanadium valence states varying mainly between V3+/V4+, i.e., Na7V3(P2O7)4 ↔ Na4V3(P2O7)4 [12][53][54]. While a similar mixed-phosphate Na7V4(P2O7)4PO4 shares similar electrochemical mechanisms (two-phase transformation of crystal structure during charge/discharge and between V3+/V4+, and the presence of an intermediate phase endowing it with better kinetics by reducing the lattice mismatch energy to overcome the phase boundary migration), i.e., Na7V4(P2O7)4PO4 ↔ Na5V4(P2O7)4PO4 ↔ Na3V4(P2O7)4PO4, although the voltage plateaus a little bit lower (3.88 V vs. Na+/Na, to be specific, 3.87 and 3.89 V for V4+/V3.5+ and V3.5+/V3+, respectively) and the capacity somewhat higher [55][56].
The fluorophosphates such as Na3V2(PO4)2F3 via the substitution of PO43− with F show even higher operating voltages based on the electrochemical mechanism of V3+/V4+ redox couple; as a member of the solid solution (oxy)fluorophosphate Na3V2(PO4)2F3−2yO2y, an increased oxygen content renders it with a slightly lower operating voltage (average ~0.1 V when y = 1, via an electrochemical mechanism of V4+/V5+ redox couple), while the capacity increases slightly (ca. by 10 mA h g−1), thus for a slightly higher energy density of 500 vs. 495 W h kg−1 [Na3V2(PO4)2FO2 vs. Na3V2(PO4)2F3[5][57]. The fluorophosphates with higher inductive effect by fluorine usually demonstrate a higher average voltage (ca. 0.3–0.5 V higher) than phosphates, although similar V3+/V4+ redox transitions occur [58]; furthermore, the V3+/V4+ transition may show slightly higher voltage than the equivalent V4+/V5+ redox couple in some cases (e.g., LiVPO4F vs. VOPO4 of 4.2 vs. 4.0 V). [58]. And for (oxy)fluorophosphate such as Na3V2O2(PO4)2F, no new phase appeared during the charging–discharging process; only a simple single-phase reaction of electrode material was involved, as can be revealed by the peak shift in ex situ XRD patterns, which relates to the change in lattice parameters; furthermore, the lattice parameter c shows a larger rate of change than a and b, viz., the distortion between the ab-planes (i.e., along the c-axis direction) is relatively larger than that within the ab-plane upon Na+ intercalation or deintercalation [19]. Moreover, the volume change of fluorophosphates such as Na3V2O2(PO4)2F upon full charge is merely 2.79%, which is much smaller than that of NaxV2(PO4)3 (ca. 8.1%), olivine NaxFePO4 (ca. 17.5%), and mixed-polyanion NaxFe3(PO4)2(P2O7) (ca. 5.1%) [19], also superior to Na1.5VPO4.8F0.7 synthesized by Kang and coworkers in 2013 (once the smallest volume change record, 2.9%) [59].
Some vanadium-based materials, mainly vanadium oxides [60] and vanadium oxide bronzes (VOBs), such as VO2 (e.g., nanowires) [10], layered V2O5 [61][62][63], γ-LiV2O5 [11], monoclinic LiV3O8 and doped Li1+xV3O8 (0 ≤ x ≤ 0.2) [64], usually have extraordinarily high capacities (some may be up to over 400 mA h g−1 for LIBs or close to 330 mA h g−1 for SIBs), although the voltage output is relatively low, e.g., with average voltage output (or long plateau of voltage) near or lower than 3.0 V, which are not focused on here and are cataloged into low-voltage cathode materials. And some similar VOBs (e.g., ω-Li3V2O5), however, due to the phase transition (e.g., γ′-V2O5γ-LiV2O5ζ-Li2V2O5ω-Li3V2O5 based on the reaction process), may be classified as anode materials due to the different alkali ion storage mechanisms and much lower plateaus (wholly or partly falling into the 0–2 V potential window), viz., these materials (V2O5-based) are binary materials that can both serve as cathodes and anodes [11]. For VOBs, the capacity of these cathode materials usually with mixed valence states may extend up to 200 mA h g−1 and beyond, although these materials usually demonstrate relatively low voltage output of ~2 V (an average discharge voltage vs. Li/Li+ [32], Mg/Mg2+ [65]) compared to high-voltage vanadium-based cathodes of >3.5 V; most of their electrochemical performances, especially the potential output plateaus, are somewhat similar to those of layered vanadium oxides (e.g., α-V2O5), while some specific phases after transition or their corresponding VOBs could show elevated voltage outputs [66]. And for the ammonium ions (NH4+) in the specific VOBs, although only a small quality can be reinserted into the layered crystal structure in these batteries (i.e., not fully reversible), they play a crucial role in maintaining the structural stability and improving the electrochemical performances, which is a new insight into the understanding of the intercalation mechanism for these host materials containing ammonium ions [32]. Some other vanadium oxides with high valence, such as LiVO3, are also this kind of binary electrodes with a tuned high-/low-voltage percentage via limiting the cut-off voltage [67][68]; while some other transition-metal-based vanadates with high capacities even up to over 1000 mA g−1 (e.g., TiO2-coated porous FeVO4 nanorods [69], porous MoV2O8 nanosheets [70]) as well as some LVOs (e.g., Li3VO4 [71][72]) are typical anode materials with major capacity in the potential window of 0–2 V (or even further, 1.0–1.5 V), which are not focused on here. Vanadyl phosphates (polymorphs of VOPO4) with layered or tunnel structures can also work as high-voltage cathode materials for alkali-metal batteries; however, initial intercalation/insertion (e.g., prelithiation, presodiation, prepotassiation) is needed for rechargeable alkali-metal-ion batteries [73]. Polyanions such as NaV3(PO4)3 are typical anode materials for SIBs with high stability [74].

2.4. Battery Assembly and Recycling

Commercial LIBs usually use graphite-based anodes, while for SIBs with larger Na+ ions, the hard carbon is in fact the best choice; PIBs, however, could also adopt graphite-based materials such as synthetic graphite as anodes due to the lower electrochemical potential of K compared to Na (−2.93 vs. −2.71, SHE) and the formation of an intermediate compound (KC8) during the stage I graphite intercalation, which offer some special advantages over SIBs [75][76]. That is, the V-based SIBs could only choose the hard carbon-based anode, while the V-based LIBs and PIBs could use the hard carbon or conventional graphite-based anodes [77]. And these carbon- or alloy-based full cells possess a higher output voltage than their symmetric counterparts with NASICON-based anodes or other types of anodes such as transition-metal chalcogenides [37][46][50][78]. It should also be mentioned that the V3+/V2+ transition will endow these vanadium-based compounds with a low potential of ca. 1.4–1.8 V, thus making them a novel class of anode candidates, which could be configured with the higher-potential transitions of V3+/V4+ and/or V4+/V5+ (ca. 3.4–4.2 V) for symmetric batteries [6][7].
Electrolyte composition also has a great impact on the performance of cathodes. Taking Na3(VOPO4)2F as an example, the ternary electrolyte compositions of EC/PC/DG (2:2:1) or EC/DEC/DG (2:2:1) + 1.0 M NaClO4 can endow them with appreciably high specific capacity (i.e., 105 and 100 mA h g−1, respectively, at 0.1C) and high cycling stability, viz., EC/PC shows better performance than EC/DEC, and the addition of diglyme (DG) can further enhance their performances by constructing a more stabilized electrode–electrolyte interface (i.e., solid electrolyte interface, SEI) [79].
Due to the increasing environmental concerns about the large-scale application of rechargeable batteries, sustainability, recyclability, and making the best of batteries should be initially considered from the design stage to the end of life. For the electrode design, the bipolar electrode structure (e.g., using Na3V2(PO4)3 as a cathode material) with Al foil as the shared current collector will enhance the energy/power densities and long-term cycling stability compared to the traditional design with Cu and Al as unipolar electrodes. Also, most of the solid components can be recycled, e.g., >98.0% on average, with ~100% Na3V2(PO4)3 and ~99.1% elemental Al, and these recycled materials could be reutilized to produce new materials with almost undiminished performances [80][81].

This entry is adapted from the peer-reviewed paper 10.3390/ma17030587

References

  1. Xu, S.; Yang, Y.; Tang, F.; Yao, Y.; Lv, X.; Liu, L.; Xu, C.; Feng, Y.; Rui, X.; Yu, Y. Vanadium fluorophosphates: Advanced cathode materials for next-generation secondary batteries. Mater. Horiz. 2023, 10, 1901–1923.
  2. Zhang, S.; Tan, H.; Rui, X.; Yu, Y. Vanadium-based materials: Next generation electrodes powering the battery revolution? Acc. Chem. Res. 2020, 53, 1660–1671.
  3. Xu, X.; Xiong, F.; Meng, J.; Wang, X.; Niu, C.; An, Q.; Mai, L. Vanadium-based nanomaterials: A promising family for emerging metal-ion batteries. Adv. Funct. Mater. 2020, 30, 1904398.
  4. Ni, W.; Shi, L.-Y. Rich structural chemistry of metal phosphates/phosphonates for emerging applications: V, Ti-containing materials. In Metal Phosphates and Phosphonates: Fundamental to Advanced Emerging Applications; Gupta, R.K., Ed.; Springer International Publishing: Cham, Switzerland, 2023; pp. 41–59.
  5. Bianchini, M.; Xiao, P.; Wang, Y.; Ceder, G. Additional sodium insertion into polyanionic cathodes for higher-energy Na-ion batteries. Adv. Energy Mater. 2017, 7, 1700514.
  6. Chen, G.; Huang, Q.; Wu, T.; Lu, L. Polyanion sodium vanadium phosphate for next generation of sodium-ion batteries—A review. Adv. Funct. Mater. 2020, 30, 2001289.
  7. Sharma, L.; Adiga, S.P.; Alshareef, H.N.; Barpanda, P. Fluorophosphates: Next generation cathode materials for rechargeable batteries. Adv. Energy Mater. 2020, 10, 2001449.
  8. Jiang, Y.; Yao, Y.; Shi, J.; Zeng, L.; Gu, L.; Yu, Y. One-dimensional Na3V2(PO4)3/C nanowires as cathode materials for long-life and high rate Na-ion batteries. ChemNanoMat 2016, 2, 726–731.
  9. Cheng, Y.; Xia, Y.; Chen, Y.; Liu, Q.; Ge, T.; Xu, L.; Mai, L. Vanadium-based nanowires for sodium-ion batteries. Nanotechnology 2019, 30, 192001.
  10. Balogun, M.-S.; Luo, Y.; Lyu, F.; Wang, F.; Yang, H.; Li, H.; Liang, C.; Huang, M.; Huang, Y.; Tong, Y. Carbon quantum dot surface-engineered VO2 interwoven nanowires: A flexible cathode material for lithium and sodium ion batteries. ACS Appl. Mater. Interfaces 2016, 8, 9733–9744.
  11. Ma, J.; Chen, Y.; Zhang, Y.; Song, T.; Wang, X.; Wu, X.; Law, M.-K.; Long, B. In situ K doped γ-LiV2O5 as long-life anode and cathode for lithium ion battery. ACS Appl. Energy Mater. 2022, 5, 10402–10408.
  12. Kovrugin, V.M.; Chotard, J.-N.; Fauth, F.; Masquelier, C. Na7V3(P2O7)4 as a high voltage electrode material for Na-ion batteries: Crystal structure and mechanism of Na+ extraction/insertion by operando X-ray diffraction. J. Mater. Chem. A 2020, 8, 21110–21121.
  13. Sørensen, D.R.; Mathiesen, J.K.; Ravnsbæk, D.B. Dynamic charge-discharge phase transitions in Li3V2(PO4)3 cathodes. J. Power Sources 2018, 396, 437–443.
  14. Guo, J.Z.; Wang, P.F.; Wu, X.L.; Zhang, X.H.; Yan, Q.; Chen, H.; Zhang, J.P.; Guo, Y.G. High-energy/power and low-temperature cathode for sodium-ion batteries: In situ XRD study and superior full-cell performance. Adv. Mater. 2017, 29, 1701968.
  15. Su, B.; Wu, S.; Liang, H.; Zhou, W.; Liu, J.; Goonetilleke, D.; Sharma, N.; Sit, P.H.L.; Zhang, W.; Yu, D.Y.W. High-performance NaVO3 with mixed cationic and anionic redox reactions for Na-ion battery applications. Chem. Mater. 2020, 32, 8836–8844.
  16. Sijuade, A.A.; Eze, V.O.; Arnett, N.Y.; Okoli, O.I. Vanadium MXenes materials for next-generation energy storage devices. Nanotechnology 2023, 34, 252001.
  17. Qin, H.; Liang, S.; Chen, L.; Li, Y.; Luo, Z.; Chen, S. Recent advances in vanadium-based nanomaterials and their composites for supercapacitors. Sustain. Energy Fuels 2020, 4, 4902–4933.
  18. Wei, Q.; DeBlock, R.H.; Butts, D.M.; Choi, C.; Dunn, B. Pseudocapacitive vanadium-based materials toward high-rate sodium-ion storage. Energy Environ. Mater. 2020, 3, 221–234.
  19. Peng, M.; Li, B.; Yan, H.; Zhang, D.; Wang, X.; Xia, D.; Guo, G. Ruthenium-oxide-coated sodium vanadium fluorophosphate nanowires as high-power cathode materials for sodium-ion batteries. Angew. Chem. Int. Ed. 2015, 54, 6452–6456.
  20. Wu, M.; Ni, W.; Hu, J.; Ma, J. NASICON-structured NaTi2(PO4)3 for sustainable energy storage. Nano-Micro Lett. 2019, 11, 44.
  21. Wei, F.; Zhang, Q.; Zhang, P.; Tian, W.; Dai, K.; Zhang, L.; Mao, J.; Shao, G. Review—Research progress on layered transition metal oxide cathode materials for sodium ion batteries. J. Electrochem. Soc. 2021, 168, 050524.
  22. Wang, Q.; Chu, S.; Guo, S. Progress on multiphase layered transition metal oxide cathodes of sodium ion batteries. Chin. Chem. Lett. 2020, 31, 2167–2176.
  23. Guo, S.; Sun, Y.; Yi, J.; Zhu, K.; Liu, P.; Zhu, Y.; Zhu, G.-Z.; Chen, M.; Ishida, M.; Zhou, H. Understanding sodium-ion diffusion in layered P2 and P3 oxides via experiments and first-principles calculations: A bridge between crystal structure and electrochemical performance. NPG Asia Mater. 2016, 8, e266.
  24. Wang, X.; Xiao, Z.; Han, K.; Zhang, X.; Liu, Z.; Yang, C.; Meng, J.; Li, M.; Huang, M.; Wei, X.; et al. Advances in fine structure optimizations of layered transition-metal oxide cathodes for potassium-ion batteries. Adv. Energy Mater. 2023, 13, 2202861.
  25. 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, 21387–21407.
  26. Wu, J.; Xu, Y.; Sun, X.; Wang, C.; Zhang, B.; Zhao, J. The multiple effects of potassium doping on LiVPO4F/C composite cathode material for lithium ion batteries. J. Power Sources 2018, 396, 155–163.
  27. Xu, C.; Zhao, J.; Wang, E.; Liu, X.; Shen, X.; Rong, X.; Zheng, Q.; Ren, G.; Zhang, N.; Liu, X.; et al. A Novel NASICON-typed Na4VMn0.5Fe0.5(PO4)3 cathode for high-performance na-ion batteries. Adv. Energy Mater. 2021, 11, 2100729.
  28. Xu, C.; Xiao, R.; Zhao, J.; Ding, F.; Yang, Y.; Rong, X.; Guo, X.; Yang, C.; Liu, H.; Zhong, B.; et al. Mn-rich phosphate cathodes for Na-ion batteries with superior rate performance. ACS Energy Lett. 2022, 7, 97–107.
  29. Zhou, H.; Cao, Z.; Zhou, Y.; Li, J.; Ling, Z.; Fang, G.; Liang, S.; Cao, X. Unlocking rapid and robust sodium storage of fluorophosphate cathode via multivalent anion substitution. Nano Energy 2023, 114, 108604.
  30. Li, P.; Gao, M.; Wang, D.; Li, Z.; Liu, Y.; Liu, X.; Li, H.; Sun, Y.; Liu, Y.; Niu, X.; et al. Optimizing vanadium redox reaction in Na3V2(PO4)3 cathodes for sodium-ion batteries by the synergistic effect of additional electrons from heteroatoms. ACS Appl. Mater. Interfaces 2023, 15, 9475–9485.
  31. Esparcia, E.; Joo, J.; Lee, J. Vanadium oxide bronzes as cathode active materials for non-lithium-based batteries. CrystEngComm 2021, 23, 5267–5283.
  32. Heo, J.W.; Bu, H.; Hyoung, J.; Hong, S.-T. Ammonium vanadium bronze, (NH4)2V7O16, as a new lithium intercalation host material. Inorg. Chem. 2020, 59, 4320–4327.
  33. Dong, Y.; Xu, J.; Chen, M.; Guo, Y.; Zhou, G.; Li, N.; Zhou, S.; Wong, C.-P. Self-assembled NaV6O15 flower-like microstructures for high-capacity and long-life sodium-ion battery cathode. Nano Energy 2020, 68, 104357.
  34. Assadi, M.H.N.; Okubo, M.; Yamada, A.; Tateyama, Y. Possible high-potential ilmenite type Na1MO3 (M = V–Ni) cathodes realized by dominant oxygen redox reaction. Phys. Rev. Mater. 2020, 4, 015401.
  35. Kuang, Q.; Zhao, Y.; Dong, Y.; Fan, Q.; Lin, X.; Liu, X. A comparative study of Li8NaV3(P2O7)3(PO4)2 and Li9V3(P2O7)3(PO4)2: Synthesis, structure and electrochemical properties. J. Power Sources 2016, 306, 337–346.
  36. Jin, T.; Liu, Y.; Li, Y.; Cao, K.; Wang, X.; Jiao, L. Electrospun NaVPO4F/C nanofibers as self-standing cathode material for ultralong cycle life Na-ion batteries. Adv. Energy Mater. 2017, 7, 1700087.
  37. Duan, W.; Zhu, Z.; Li, H.; Hu, Z.; Zhang, K.; Cheng, F.; Chen, J. Na3V2(PO4)3@C core–shell nanocomposites for rechargeable sodium-ion batteries. J. Mater. Chem. A 2014, 2, 8668–8675.
  38. Ren, W.; Zheng, Z.; Xu, C.; Niu, C.; Wei, Q.; An, Q.; Zhao, K.; Yan, M.; Qin, M.; Mai, L. Self-sacrificed synthesis of three-dimensional Na3V2(PO4)3 nanofiber network for high-rate sodium–ion full batteries. Nano Energy 2016, 25, 145–153.
  39. Ni, W. Green, sustainable and massive synthesis of sodium vanadate nanowires toward industrialization. Prog. Progress. Nat. Sci. Mater. Int. 2024, 34.
  40. Pan, A.; Choi, D.; Zhang, J.-G.; Liang, S.; Cao, G.; Nie, Z.; Arey, B.W.; Liu, J. High-rate cathodes based on Li3V2(PO4)3 nanobelts prepared via surfactant-assisted fabrication. J. Power Sources 2011, 196, 3646–3649.
  41. Yang, Y.; Xu, S.; Zhu, W.; Xu, C.; Rui, X. Electrospinning polyanionic materials for high-Rate Na storage. Batter. Supercaps 2023, 6, e202300271.
  42. Subramanian, Y.; Oh, W.; Choi, W.; Lee, H.; Jeong, M.; Thangavel, R.; Yoon, W.-S. Optimizing high voltage Na3V2(PO4)2F3 cathode for achieving high rate sodium-ion batteries with long cycle life. Chem. Eng. J. 2021, 403, 126291.
  43. Shen, X.; Zhou, Q.; Han, M.; Qi, X.; Li, B.; Zhang, Q.; Zhao, J.; Yang, C.; Liu, H.; Hu, Y.-S. Rapid mechanochemical synthesis of polyanionic cathode with improved electrochemical performance for Na-ion batteries. Nat. Commun. 2021, 12, 2848.
  44. Wang, J.; Wang, Z.; Ni, J.; Li, L. Electrospinning for flexible sodium-ion batteries. Energy Storage Mater. 2022, 45, 704–719.
  45. Rui, X.; Sun, W.; Wu, C.; Yu, Y.; Yan, Q. An advanced sodium-ion battery composed of carbon coated Na3V2(PO4)3 in a porous graphene network. Adv. Mater. 2015, 27, 6670–6676.
  46. Fang, Y.; Xiao, L.; Ai, X.; Cao, Y.; Yang, H. Hierarchical carbon framework wrapped Na3V2(PO4)3 as a superior high-rate and extended lifespan cathode for sodium-ion batteries. Adv. Mater. 2015, 27, 5895–5900.
  47. Tertov, I.V.; Drozhzhin, O.A.; Alekseeva, A.M.; Kirsanova, M.A.; Mironov, A.V.; Abakumov, A.M.; Antipov, E.V. β-LiVP2O7 as a positive electrode material for Li-ion batteries. Electrochim. Acta 2021, 389, 138759.
  48. Drozhzhin, O.A.; Tertov, I.V.; Alekseeva, A.M.; Aksyonov, D.A.; Stevenson, K.J.; Abakumov, A.M.; Antipov, E.V. β-NaVP2O7 as a superior electrode material for Na-ion batteries. Chem. Mater. 2019, 31, 7463–7469.
  49. Lin, X.; Zhao, Y.; Yan, D.; Han, W.; Kuang, Q.; Wen, M. Layered Li-rich vanadium phosphate Li9V3(P2O7)3(PO4)2: Cathode and anode materials for lithium-ion batteries. Electrochim. Acta 2016, 191, 207–214.
  50. Saravanan, K.; Mason, C.W.; Rudola, A.; Wong, K.H.; Balaya, P. The first report on excellent cycling stability and superior rate capability of Na3V2(PO4)3 for sodium ion batteries. Adv. Energy Mater. 2013, 3, 444–450.
  51. Min, X.; Huo, H.; Li, R.; Zhou, J.; Hu, Y.; Dai, C. Cycling stability of Li3V2(PO4)3/C cathode in a broad electrochemical window. J. Electroanal. Chem. 2016, 774, 76–82.
  52. Jia, Y.; Wu, Y.; Li, L.; Song, L.; Gao, J. Monoclinic Na2VOP2O7: A 4V-class cost-effective cathode for sodium-ion batteries. Mater. Today Phys. 2023, 33, 101038.
  53. Kim, J.; Park, I.; Kim, H.; Park, K.-Y.; Park, Y.-U.; Kang, K. Tailoring a new 4V-class cathode material for Na-ion batteries. Adv. Energy Mater. 2016, 6, 1502147.
  54. Deng, C.; Zhang, S.; Zhao, B. First exploration of ultrafine Na7V3(P2O7)4 as a high-potential cathode material for sodium-ion battery. Energy Storage Mater. 2016, 4, 71–78.
  55. Lim, S.Y.; Kim, H.; Chung, J.; Lee, J.H.; Kim, B.G.; Choi, J.-J.; Chung, K.Y.; Cho, W.; Kim, S.-J.; Goddard, W.A.; et al. Role of intermediate phase for stable cycling of Na7V4(P2O7)4PO4 in sodium ion battery. Proc. Natl. Acad. Sci. USA 2014, 111, 599–604.
  56. Deng, C.; Zhang, S. 1D Nanostructured Na7V4(P2O7)4(PO4) as high-potential and superior-performance cathode material for sodium-ion batteries. ACS Appl. Mater. Interfaces 2014, 6, 9111–9117.
  57. Park, Y.-U.; Seo, D.-H.; Kim, H.; Kim, J.; Lee, S.; Kim, B.; Kang, K. A family of high-performance cathode materials for Na-ion batteries, Na3(VO1−xPO4)2F1+2x (0 ≤ x ≤ 1): Combined first-principles and experimental study. Adv. Funct. Mater. 2014, 24, 4603–4614.
  58. Barker, J.; Saidi, M.Y.; Swoyer, J.L. Electrochemical insertion properties of the novel lithium vanadium fluorophosphate, LiVPO4 F. J. Electrochem. Soc. 2003, 150, A1394.
  59. Park, Y.-U.; Seo, D.-H.; Kwon, H.-S.; Kim, B.; Kim, J.; Kim, H.; Kim, I.; Yoo, H.-I.; Kang, K. A new high-energy cathode for a Na-ion battery with ultrahigh stability. J. Am. Chem. Soc. 2013, 135, 13870–13878.
  60. Zhang, X.; Zhang, Z.; Xu, S.; Xu, C.; Rui, X. Advanced vanadium oxides for sodium-ion batteries. Adv. Funct. Mater. 2023, 33, 2306055.
  61. Niu, X.; Li, N.; Chen, Y.; Zhang, J.; Yang, Y.; Tan, L.; Wu, J.; Guo, L.; Zhu, Y. Structure evolution of V2O5 as electrode materials for metal-ion batteries. Batter. Supercaps 2023, 6, e202300238.
  62. Yao, J.; Li, Y.; Massé, R.C.; Uchaker, E.; Cao, G. Revitalized interest in vanadium pentoxide as cathode material for lithium-ion batteries and beyond. Energy Storage Mater. 2018, 11, 205–259.
  63. Liu, Q.; Li, Z.-F.; Liu, Y.; Zhang, H.; Ren, Y.; Sun, C.-J.; Lu, W.; Zhou, Y.; Stanciu, L.; Stach, E.A.; et al. Graphene-modified nanostructured vanadium pentoxide hybrids with extraordinary electrochemical performance for Li-ion batteries. Nat. Commun. 2015, 6, 6127.
  64. Zhu, L.; Ge, P.; Xie, L.; Miao, Y.; Cao, X. Doped-Li1+xV3O8 as cathode materials for lithium-ion batteries: A mini review. Electrochem. Commun. 2020, 115, 106722.
  65. Esparcia, E.A., Jr.; Chae, M.S.; Ocon, J.D.; Hong, S.-T. Ammonium vanadium bronze (NH4V4O10) as a high-capacity cathode material for nonaqueous magnesium-ion batteries. Chem. Mater. 2018, 30, 3690–3696.
  66. Christensen, C.K.; Sørensen, D.R.; Hvam, J.; Ravnsbæk, D.B. Structural evolution of disordered LixV2O5 bronzes in V2O5 cathodes for Li-ion batteries. Chem. Mater. 2019, 31, 512–520.
  67. Yang, D.; Zhang, D.; Wu, H.; Pei, C.; Xiao, T.; Ma, H.; Ni, S. Construction of a LiVO3/C core–shell structure for high-rate lithium storage. New J. Chem. 2023, 47, 1508–1516.
  68. Fu, X.; Pu, X.; Wang, H.; Zhao, D.; Liu, G.; Zhao, D.; Chen, Z. Understanding capacity fading of the LiVO3 cathode material by limiting the cutoff voltage. Phys. Chem. Chem. Phys. 2019, 21, 7009–7015.
  69. Zheng, H.; Yang, Y.; Liu, X.; Guo, Z.; Feng, C. Controllable synthesis of FeVO4@TiO2 nanostructures as anode for lithium ion battery. J. Nanopart. Res. 2017, 19, 243.
  70. Zheng, H.; Zhang, Q.; Gao, H.; Sun, W.; Zhao, H.; Feng, C.; Mao, J.; Guo, Z. Synthesis of porous MoV2O8 nanosheets as anode material for superior lithium storage. Energy Storage Mater. 2019, 22, 128–137.
  71. Zhang, C.; Song, H.; Liu, C.; Liu, Y.; Zhang, C.; Nan, X.; Cao, G. Fast and reversible Li ion insertion in carbon-encapsulated Li3VO4 as anode for lithium-ion battery. Adv. Funct. Mater. 2015, 25, 3497–3504.
  72. Liao, C.; Zhang, Q.; Zhai, T.; Li, H.; Zhou, H. Development and perspective of the insertion anode Li3VO4 for lithium-ion batteries. Energy Storage Mater. 2017, 7, 17–31.
  73. Shi, H.-Y.; Jia, Z.; Wu, W.; Zhang, X.; Liu, X.-X.; Sun, X. The development of vanadyl phosphate cathode materials for energy storage systems: A review. Chem. Eur. J. 2020, 26, 8190–8204.
  74. Hu, P.; Wang, X.; Ma, J.; Zhang, Z.; He, J.; Wang, X.; Shi, S.; Cui, G.; Chen, L. NaV3(PO4)3/C nanocomposite as novel anode material for Na-ion batteries with high stability. Nano Energy 2016, 26, 382–391.
  75. Lin, X.; Huang, J.; Tan, H.; Huang, J.; Zhang, B. K3V2(PO4)2F3 as a robust cathode for potassium-ion batteries. Energy Storage Mater. 2019, 16, 97–101.
  76. Beltrop, K.; Beuker, S.; Heckmann, A.; Winter, M.; Placke, T. Alternative electrochemical energy storage: Potassium-based dual-graphite batteries. Energy Environ. Sci. 2017, 10, 2090–2094.
  77. Xu, Y.; Titirici, M.; Chen, J.; Cora, F.; Cullen, P.L.; Edge, J.S.; Fan, K.; Fan, L.; Feng, J.; Hosaka, T.; et al. 2023 roadmap for potassium-ion batteries. J. Phys. Energy 2023, 5, 021502.
  78. Wang, Y.-Y.; Hou, B.-H.; Guo, J.-Z.; Ning, Q.-L.; Pang, W.-L.; Wang, J.; Lü, C.-L.; Wu, X.-L. An ultralong lifespan and low-temperature workable sodium-ion full battery for stationary energy storage. Adv. Energy Mater. 2018, 8, 1703252.
  79. Krishna, M.M.; Jayasree, G.; Mallu, R.R.; Basak, P. Impact of electrolyte composition on the performance of Na3(VOPO4)2F Cathodes: A Comprehensive Case Study. ACS Appl. Energy Mater. 2023, 6, 9287–9299.
  80. Liu, T.; Zhang, Y.; Chen, C.; Lin, Z.; Zhang, S.; Lu, J. Sustainability-inspired cell design for a fully recyclable sodium ion battery. Nat. Commun. 2019, 10, 1965.
  81. Samarin, A.S.; Ivanov, A.V.; Fedotov, S.S. Toward efficient recycling of vanadium phosphate-based sodium-ion batteries: A review. Clean Technol. 2023, 5, 881–900.
More
This entry is offline, you can click here to edit this entry!
Video Production Service