Next Article in Journal / Special Issue
Chlorine Adsorption on TiO2(110)/Water Interface: Nonadiabatic Molecular Dynamics Simulations for Photocatalytic Water Splitting
Previous Article in Journal
Acknowledgment to the Reviewers of Electronic Materials in 2022
Previous Article in Special Issue
Stress-Relaxed AlN-Buffer-Oriented GaN-Nano-Obelisks-Based High-Performance UV Photodetector
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Research Process of Carbon Engineering on Na3V2(PO4)3 for Sodium-Ion Battery Cathodes: A Mini Review

Ningxia Key Laboratory of Photovoltaic Materials, School of Materials and New Energy, Ningxia University, Yinchuan 750021, China
*
Author to whom correspondence should be addressed.
Electron. Mater. 2023, 4(1), 17-32; https://doi.org/10.3390/electronicmat4010003
Submission received: 21 November 2022 / Revised: 20 January 2023 / Accepted: 29 January 2023 / Published: 31 January 2023
(This article belongs to the Special Issue Feature Papers of Electronic Materials II)

Abstract

:
Owing to the 3D open framework, excellent structural stability, and high ionic conductivity, NASICON-type compounds are extensively employed as promising cathode materials for sodium-ion batteries (SIBs). Being one of the representative NASICON-type compounds, the Na3V2(PO4)3 delivers high theoretical capacity with an operating voltage exceeding 3.3 V, enabling it to be a good candidate for SIBs. Unfortunately, the Na3V2(PO4)3 suffers from low electronic conductivity. In this work, we briefly review the recent research progress on novel carbon engineering strategies to enhance the electronic conductivity of Na3V2(PO4)3. Moreover, we will point out the issues relating to the development of NASICON cathode materials and put forward some suggestions.

1. Introduction

Lithium-ion batteries (LIBs) are widely used in high-energy storage fields, such as electric vehicles, communications, and smart electronics [1,2,3,4]. However, the development of LIBs is threatened by the limited lithium resources where the earth’s crust content of lithium is only 0.0065% [5,6,7]. Therefore, it is crucial to develop alternatives to replace LIBs in the future. Compared to lithium, the sodium reserve is abundant in the earth’s crust (2.8%) and sea (10.77g/kg) [8], and Na exhibits similar properties as Li since they all settle in group IA of the periodic tables. As realized in Table 1, Na also has a low standard electrode potential of −2.71 V, which is just slightly higher than Li (−3.04 V) [9,10,11,12,13,14]. As a result, SIBs have similar working principles to LIBs in that Na+ can reversibly shuttle between cathode and anode, causing a voltage potential difference that leads to the mutual conversion of electrical and chemical energy. However, the large ionic radius (1.02 Å) and atomic mass (22.99 g/mol) of Na+ result in low theoretical capacity, high resistance, and sluggish transport kinetics during the Na+ insertion/de-insertion [15,16,17,18]. Thus, some traditional electrode materials applied for LIBs are not suitable in SIBs directly, such as the extensively used LiCoO2 cathode and the graphite anode. Therefore, it is significant to explore SIB electrode materials with exceptional Na+ insertion/de-insertion kinetics and Na+ storage performance. Associating with electrochemical batteries, cathode materials as an essential component of batteries, make great contributions to strengthening electrochemical performance. This is because the inherent crystal structure of the cathode materials determines the electrochemical behavior of SIBs, including redox potential, the number of electron transfers, and structural stability. The main cathode materials adapted to SIBs can be classified into three types: (1) stratified material [19,20,21,22], (2) polyanionic compounds [23,24,25,26], and (3) macromolecule polymer material [27,28,29,30,31]. Further, Table 2 shows the performance of different cathode materials for SIBs.
Polyanionic compounds have been regarded as one of the most attractive cathode materials due to their Na+ superionic conductor (NASICON) structure and adjustable redox potential [46,47,48,49,50]. The general formula of NASICON-type compounds is described as AxM2(XO4)3 (A = Na, Li, Fe, etc., M = V, Ti, Fe, Tr or Nb, etc., X = P or S and x = 0–4). The NASICON-type AxM2(XO4)3 is constructed by corner-shared MO6 and polyhedral XO4 groups, which provide convenient Na+ diffusion channels [51]. The open 3D framework and large pore channels make the polyanionic compounds possess high Na+ conductivity and long-term structure stability [52,53,54,55,56,57]. Among various NASICON-type AxM2(XO4)3, the Na3V2(PO4)3 (NVP) cathode offers a high Na+ storage capacity of 117 mAh/g with high structural stability. However, the separation of TiO6 octahedrons leads to poor electronic conductivity, which further influence rate capability and cyclic stability [58,59,60,61,62,63]. In order to overcome these problems, many advanced strategies have been proposed. The typical approaches involve particle nano-sizing, doping, and coating with carbons, including graphene, carbon black, and carbon nanotubes (CNTs), etc. [64,65,66,67,68,69].
In this review, we initially analyze the structure−performance relationship of NVP for SIBs and then briefly summarize the research process of the carbon coating approaches to modify the NVP. Significantly, the modifications that can improve the conductivity of NVP by carbon coating strategies are emphatically introduced. Moreover, the challenges and future development direction of the NASICON-type cathode have been prospected. It is expected that this work will be a meaningful reference to guide the design of high-performance NVP cathodes for SIBs.

2. The Structure and Properties of NVP

Vanadium has rich oxidation valence states, which contain V5+, V4+, V3+, and V2+. Theoretically, the NVP cathode reveals a high capacity of 117 mAh/g because of the conversion of V3+/V4+ [70,71]. Meanwhile, its highly open framework enables the creation of 3D diffusion channels and large migration gaps for Na+, and the diffusion coefficient of Na+ can reach 10−10 cm2/s [72,73]. Figure 1 shows the crystal structure of Na3V2(PO4)3 generated by VESTA with the monoclinic lattice, and the space group is P21/c. Besides, lattice parameters are a = 8.87 Å, b = 9.01 Å, c = 15.44 Å and α = 55.326°, β = 90° and γ = 90°. The structure is composed of VO6 octahedrons and PO4 tetrahedrons by sharing an O atom at the vertex. At the same time, two free Na+ occupy the two oxidation state channels in the lattice. This structure promises excellent structural stability for NVP. For example, Choi et al. [74] studied the thermal stability of the NVP cathode by using TGA-DSC and in situ XRD at various temperatures. Basically, the NVP exhibits no endothermic/exothermic peak and prominent thermal stability even at 450 °C under a deoxidation state, which is ascribed to the high stability of the P-O bond. Moreover, the NVP illuminates the typical redox potentials at 3.4 V vs. Na/Na+, which is higher than that of the majority of cathode materials [75,76,77]. When employed as a cathode, the cell reaction can be described as:
Cathode :   Na 3 V 2 3 + PO 4 3 Na 3 x V x 4 + PO 4 3 + xNa + + xe
In the charging process, two Na+ are expelled from the cathode to the anode. Meanwhile, V3+ is oxidized to V4+ to form NaV2(PO4)3, accompanied by the transfer of two electrons. On the contrary, V4+ is reduced to V3+, and Na+ is inserted into the cathode to form NVP.

3. The Modification Methods to Improve the Conductivity of NVP

Despite the fact that the NVP demonstrates an advance as a cathode for SIBs, the conductivity of NVP is still concerning. Carbon coating engineering is often considered to be one of the most effective methods to improve the electronic conductivity of base materials. It has many advantages, such as low cost, simple operation, and flexibility.

3.1. Graphene

Graphene is a 2D carbon allotrope constituted by sp2 hybridization. The remaining p orbital electrons in the plane form a large π bond to move freely over graphene, resulting in fast electron mobility and high electrical conductivity. The charge carrier mobility and thermal conductivity of graphene exceed 2 × 105 cm2/Vs and 3000 W/Mk at an electron density of 2 × 1011 cm−2, respectively. As a result, graphene is employed as a conductive component to composite with NVP by either coating or doping. For example, Rui et al. [78] proposed a 3D NVP-based hybrid cathode, which is composed of NVP@amorphous carbon wrapped by reduced graphene oxide nanosheets (NVP@C@rGO) to achieve efficient charge transfer, as shown in Figure 2a,b. Essentially, the aqueous NVP precursor is mixed with poly (vinylpyrrolidone) (PVP) to construct a nanoporous carbon matrix on NVP nanoparticles. Then, the GO suspension was added to the above solution and followed by performing freeze-drying and heating. In Figure 2c,d, the NVP@C@rGO revealed prominent rate performance; that is, the discharge capacity remained at about 115 mAh/g when the current density varies from 1 to 20 C. Figure 2e shows that the charge transfer resistance (Rct) of NVP@C@rGO was about 258 Ω, which was much lower than that of NVP@C (639 Ω). Chang et al. [79] synthesized a graphene-bound, binder-free NVP film electrode by simple vacuum-assisted filtration of the NVP and graphene. Figure 2f shows the SEM image of the as-prepared graphene-bound NVP film. Obviously, the 100–300 nm NVP particles are evenly bound to the graphene. Compared with pristine NVP particles, the graphene not only increased the conductivity of the whole electrode but also suppressed the agglomeration of NVP particles. As a result, it exhibited excellent performance under a high rate (Figure 2g). Further, in Figure 2h, the Rct of graphene-bound NVP was significantly smaller than that of pristine NVP, which again confirms that the GO benefits alleviate the electronic conductivity of NVP.
Besides, Xu et al. [80] prepared NVP/reduced graphene oxide hollow spheres (NVP/rGO HSs) by spray drying and subsequent pyrolysis for Na+ storage (Figure 3a). The SEM and TEM images of NVP/rGO HSs are displayed in Figure 3b,c. The graphene layer was surrounded by a large number of NVP particles, which together formed a hollow sphere (~6 μm). The spherical hollow structure can make full contact between the electrode and electrolyte, and the rGO layer acted as a high-speed electron transferring bridge. Therefore, NVP/rGO HSs exhibited excellent electrochemical performance as well as remarkable rate performance, i.e., the specific discharge capacity was 107.5 mAh/g at 10 C. Further, Xu et al. [81] designed a straightforward and self-assembly approach to fabricate layer-by-layer structured NVP@reduced graphene oxide (NVP@rGO) nanocomposites. In NVP@rGO, the rGO was responsible for offering rapid electron transport. Thus, the NVP@rGO cathode presents superior rate capability (73 mAh/g at 100 C and 41 mAh/g at 200 C).

3.2. Carbon Nanotubes

Besides graphene, CNTs are also chosen as conductive blocks to couple with NVP. CNTs can be viewed as coiled from graphene sheets, and the p orbital electrons on CNTs provide more π bonds to produce a strong conjugation effect. As a result, high electrical conductivity is associated with CNTs. For example, Shen et al. [82] developed a simple sol-gel approach to prepare double nano-carbon modified NVP (NVP@C+N@CNTs) to improve the electrochemical characteristic of SIBs, and the corresponding schematic structures of NVP@C+N@CNTs are drawn in Figure 4a. In Figure 4b,c, associating with the TEM image of NVP@C+N@CNTs, it is clear to observe that the NVP is perfectly covered by carbon coatings and simultaneously connected by CNTs to form a 3D network. Such a structure accelerated the electron transport between NVP nanoparticles. In Figure 4d, the specific discharge capacities of NVP@C+N@CNTs reach 94.5, 94.6, 94.5, 93.6, 93.4, 92.6, 91.5, 92.5, 92.4, and 90.1 mAh/g at 0.2, 0.5, 1, 2, 3, 5, 10, 15, 20, and 30 C, respectively. Even at 70 C, the specific discharge capacity still remained at 70 mAh/g.
Chen et al. [83] designed and synthesized a CNT-decorated NVP microsphere via spray drying and a carbothermal reduction approach. The NVP microspheres were embedded by CNTs. Meanwhile, the particles were coated by amorphous carbon layers. Figure 5a–d shows the SEM images of NVP doped with different amounts of CNTs. As the CNTs content increases, fewer NVP particles are observed. This is presumably ascribed to the fact that the CNTs limited the growth of NVP particles. Further, the CNTs and carbon layer can be apparently examined from the TEM images in Figure 5e,f. In this structure, the unique carbon network effectively promoted electron transmission. Most importantly, the incorporation of CNTs significantly enhanced the rate capability (Figure 5g). By comparison with NVP/C, the NVP/C10 illustrated a smaller Rct (225.5 Ω), which is also beneficial to improve the electrochemical properties of the composite electrode (Figure 5h). Beyond these, Sun et al. [84] prepared Na3V2(PO4)3/C@CNTs-WC by a sol-gel method. The C@CNTs formed a conductive framework that accelerated electron transport. At the same time, the introduction of W6+ created abundant cavities on the surface of the NVP. Therefore, the Na3V2(PO4)3/C@CNTs-WC delivers a high capacity of 92.6 mAh/g at a super high rate of 50 C and remains 84.5 mAh/g over 400 cycles.

3.3. Other Carbon Coatings

In any case, the carbon coating layer (CCL) can inhibit the size of NVP particles during growth. Besides, the CCL acted as a buffer layer to avoid direct contact between the electrolyte and active materials, which helped to weaken the corrosion of the electrolyte over the electrode. Moreover, the CCL is usually amorphous carbon, which has good mechanical properties and stability. For instance, Duan et al. [85] employed a hydrothermal-assisted sol-gel method to synthesize the core-shelled NVP@C. Notably, there presents a large gap between the as-prepared nano NVP@C particles, which is conducive to urging complete contact between the electrolyte and the NVP@C and thus shortening the Na+ diffusion path. When it was applied for SIBs, the nano NVP@C realized the high specific discharge capacity of 88 mAh/g at 10 C. In the work performed by Li et al. [86], citric acid was used as a carbon resource, reducing agent, and complexing agent. They proposed a method by rheological phase reaction to prepare double carbon-wrapped NVP where amorphous carbon served as a framework together with graphitized carbon, which acted as a conductive network. By this design, the thickness of amorphous carbon is about 5 nm; electrons are quickly transferred to amorphous carbon through graphitized carbon and then transferred to NVP from amorphous carbon, which effectively shortens the transmission path of electrons. Consequently, the NVP@C calcined at 800 °C (NVPC-800) illuminated the best electrochemical performance among all samples calcined at 600, 700, and 900 °C. In Figure 6a,b, relating to the TEM images, the thickness of the disordered carbon layer was ~5 nm, and the graphitized carbon can be clearly observed as well. Figure 6c,d show the cycling performance and EIS spectra of NVPC-800, verifying the prominent cycling stability and low Rct. Even the current density increased to 1 C, and the specific discharge capacity of NVPC-800 reached 99.8 mAh/g with more than 98.9% capacity retention after 500 cycles (Figure 6e).
Jiang et al. [87] synthesized carbon-coated nanosized NVP into mesoporous carbons (NVP@C@CMK-3) by a nano-coating technique, as drawn in Figure 7a. It schematically described the reversible electrochemical reactions of the NVP@C nanoconfined within a single carbon pore that the special structure ensured efficient diffusion of electrons and Na+. In Figure 7b, the NVP@C@CMK-3 reveals the high specific discharge capacity of 115 mAh/g at 110 mA/g, which is much higher than that of pure NVP (83 mAh/g). From Figure 7c, the Rct of the NVP@CMK-3 and NVP@C@CMK-3 is 272 and 207 Ω, respectively, highlighting that the double carbon coating structure facilitates the transportation of Na+ and e and meanwhile generates lower polarization [88,89].

3.4. Heteroatom Doping

Researchers have found that either anion or cation doping is in favor of improving the intrinsic conductivity of NVP, such as Ag+, N4-, etc. Basically, ion doping leads to a greater carrier concentration of the doped system over the intrinsic system, which would significantly enhance conductivity. Hong et al. [90] prepared C-Ag co-coated NVP (NVP/C-Ag) and C-coated NVP (NVP/C) by an uncomplicated solid-state method in Figure 8a. From the XRD patterns of NVP/C-Ag in Figure 8b, the typical characteristic peak of silver was captured at 38°. Meanwhile, the diffraction peak of NVP/C-Ag (at about 23.5°) was wider and lower than that of NVP/C, which is attributed to the confined growth of NVP particles due to the presence of the C-Ag layer. Figure 8c displays the TG curves of NVP/C-Ag. When the temperature exceeded 500 °C, the carbon species had completely burned out, and the residual carbon content was about 6%. Figure 8d depicts the charge/discharge curves of NVP/C-Ag at various rates. Upon the current density increases, small changes can be obtained in the specific discharge capacity, featuring excellent rate performance. In Figure 8e, Wang et al. [91] prepared NVP/N-doped carbon composites (NVP-NC) by taking nitrilotrimethylene triphosphonic acid as the multifunctional raw material. Figure 8f shows the SEM image of NVP-NC. It can be seen that NTP particles were encapsulated in an N-doped carbon matrix, which facilitates the rapid transport of Na+ and electrons. Among all samples, the NVP-NC obtained under 850 °C showed the best Na+ storage performance, and the corresponding specific discharge capacity was 94.0 mAh/g at 10 C after 600 cycles.
Jiang et al. [92] designed an (N, S) co-doped graphene (3DPGFs-NS) to accommodate the active materials. Then, the Na3V2(PO4)3@C@3DPGFs-NS were prepared by a sol-gel method, as presented in Figure 9a. From the SEM image of Figure 9b, the obvious 3D porous structure was observed. Further, the carbon layer, NVP, and graphene can be clearly differentiated from the TEM image in Figure 9c. The charge/discharge curves of NVP and Na3V2(PO4)3@C@3DPGFs-NS at 110 mA/g are shown in Figure 9d. The Na3V2(PO4)3@C@3DPGFs-NS displays a wider discharge platform, and the curve is smoother, suggesting the outstanding electrical conductivity of (N, S) co-doped 3DPGFs. The specific discharge capacity of Na3V2(PO4)3@C@3DPGFs-NS is 112 mAh/g at 1 C, which is higher than pure NVP (60 mAh/g). Further, the Na3V2(PO4)3@C@3DPGFs-NS reveals excellent cycling ability at 20 C in Figure 9e. Above all, the electrochemical performance of NVP has been improved through the carbon coating and heteroatom doping strategy in Table 3.

4. Conclusions and Outlook

NVP has demonstrated great potential in SIBs as the cathode due to its high operating voltage, high ionic conductivity, and good structural and thermal stability. In this review, we briefly analyze various carbon coating strategies, such as graphene, carbon nanotubes, and other carbon allotropes, to improve the electronic conductivity of NVP towards high-performance SIBs. The results suggest that various carbons not only ameliorate the structure of NVP but also enhance the conductivity of the whole electrode. Meanwhile, the carbon matrix can store Na+ by itself. Beyond that, heteroatom doping has also been employed to enhance the conductivity of pure NVP by increasing the carrier concentration. Despite the NVP cathode materials have developed rapidly in recent years, there are still some key issues that have to be resolved before the application of NVP on a large scale. Firstly, high-capacity NVP electrodes should be designed by involving multi-electron reactions. In this regard, the multi-scale computational models are necessarily put forward as a guide. Further, it is still a challenge to achieve high conductive NVP with excellent electrochemical properties. Fortunately, great efforts have been devoted to dope transition metal ions into NVP to improve the electronic conductivity from the inherent structure of the material, such as Mg2+ and Mn2+, etc. On the other hand, the substitution of V sites by Hf, Zr, and W deserves to be explored as well. Finally, high voltage electrolyte is essential to achieve the multi-electron transfer of NVP cathodes and thus higher energy density.

Funding

This work was supported by the project of Ningxia’s key R&D plan (2021BEE03006).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing does not apply to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. He, Z.; Jiang, Y.; Meng, W.; Zhu, J.; Liu, Y.; Dai, Y.; Wang, L.L. Advanced LiTi2(PO4)3@N-doped carbon anode for aqueous lithium ion batteries. Electrochim. Acta 2016, 222, 1491–1500. [Google Scholar] [CrossRef]
  2. Marom, R.; Amalraj, S.F.; Leifer, N.; Jacob, D.; Aurbach, D. A review of advanced and practical lithium battery materials. J. Mater. Chem. 2011, 21, 9938–9954. [Google Scholar] [CrossRef]
  3. Goodenough, J.B.; Park, K.S. The Li-ion rechargeable battery: A perspective. J. Am. Chem. Soc. 2013, 135, 1167–1176. [Google Scholar] [CrossRef] [PubMed]
  4. Li, Y.; Chen, Y.; Luo, L.; Ming, X.; Li, J.; Lei, T.; Deng, S.; Guo, J.; Zhu, J.; Chang, S. Surface in-situ reconstruction of LiNi0.8Co0.1Mn0.1O2 cathode materials interacting with antimony compounds and the electrochemical performances. J. Electroanal. Chem. 2019, 854, 113582. [Google Scholar] [CrossRef]
  5. Cao, X.; Sun, Q.; Zhu, L.; Xie, L. Na3V2(PO4)3 nanoparticles confined in functional carbon framework towards high-rate and ultralong-life sodium storage. J. Alloys Compd. 2019, 791, 296–306. [Google Scholar] [CrossRef]
  6. Yi, H.; Lin, L.; Ling, M.; Lv, Z.; Li, R.; Fu, Q.; Zhang, H.; Zheng, Q.; Li, X. Scalable and economic synthesis of high-performance Na3V2(PO4)2F3 by a solvothermal-ball-milling method. ACS Energy Lett. 2019, 4, 1565–1571. [Google Scholar] [CrossRef]
  7. Liu, T.; Zhang, Y.; Jiang, Z.; Zeng, X.; Ji, J.; Li, Z.; Gao, X.; Sun, M.; Lin, Z.; Ling, M.; et al. Exploring competitive features of stationary sodium ion batteries for electrochemical energy storage. Energy Environ. Sci. 2019, 12, 1512–1533. [Google Scholar] [CrossRef]
  8. Zhang, B.W.; Ma, K.X.; Lv, X.; Shi, K.; Wang, Y.; Nian, Z.Y.; Li, Y.H.; Wang, L.; Dai, L.; He, Z.X. Recent advances of NASICON-Na3V2(PO4)3 as cathode for sodium-ion batteries: Synthesis, modifications, and perspectives. J. Alloys Compd. 2021, 867, 159060. [Google Scholar] [CrossRef]
  9. Chagas, L.G.; Jeong, S.; Hasa, I.; Passerini, S. Ionic liquid-based electrolytes for sodium-ion batteries: Tuning properties to enhance the electrochemical performance of manganese-based layered oxide cathode. ACS Appl. Mater. Inter. 2019, 11, 22278–22289. [Google Scholar] [CrossRef]
  10. Gu, Y.; Xiao, F.; Xu, K.; Pan, X.; Li, J.; Xu, C.; Zhou, X. Outstanding electrochemical performance of sodium vanadium phosphate cathode co-modified by carbon-coating and titanium-doping for Na-ion batteries. Ceram. Int. 2019, 45, 12570–12574. [Google Scholar] [CrossRef]
  11. Zhao, J.; Gao, Y.; Liu, Q.; Meng, X.; Chen, N.; Wang, C.; Du, F.; Chen, G. High rate capability and enhanced cyclability of Na3V2(PO4)2F3 cathode by in situ coating of carbon nanofibers for sodium-ion battery applications. Chem. Eur. J. 2018, 24, 2913–2919. [Google Scholar] [CrossRef] [PubMed]
  12. Slater, M.D.; Kim, D.; Lee, E.; Johnson, C.S. Sodium-ion batteries. Adv. Funct. Mater. 2013, 23, 947–958. [Google Scholar] [CrossRef]
  13. Kim, S.W.; Seo, D.H.; Ma, X.; Ceder, G.; Kang, K. Electrode materials for rechargeable sodium-ion batteries: Potential alternatives to current lithium-ion batteries. Adv. Energy Mater. 2012, 2, 710. [Google Scholar] [CrossRef]
  14. Zhou, Y.X.; Zhang, L.L.; Yang, X.L.; Huang, Y.H.; Ding, X.K.; Ma, D.; Wang, J.Q. Synthesis of nanosheet-structured Na3V2(PO4)3/C as high-performance cathode material for sodium ion batteries using anthracite as carbon source. Ceram. Int. 2017, 43, 2333. [Google Scholar] [CrossRef]
  15. Khan, Z.; Vagin, M.; Crispin, X. Can hybrid Na–air batteries outperform nonaqueous Na–O2 batteries? Adv. Sci. 2020, 7, 1902866. [Google Scholar] [CrossRef] [Green Version]
  16. Zhao, Y.; Yang, D.; He, T.; Li, J.; Wei, L.; Wang, D.; Wang, Y.; Wang, X.; Chen, G.; Wei, Y. Vacancy engineering in VS2 nanosheets for ultrafast pseudocapacitive sodium ionstorage. Chem. Eng. J. 2021, 421, 129715. [Google Scholar] [CrossRef]
  17. Lv, Z.Q.; Ling, M.X.; Yue, M.; Li, X.F.; Song, M.M.; Zheng, Q.; Zhang, H.M. Vanadium-based polyanionic compounds as cathode materials for sodium-ion batteries: Toward high-energy and high-power applications. J. Energy Chem. 2021, 55, 361–390. [Google Scholar] [CrossRef]
  18. Chen, S.; Wu, C.; Shen, L.; Zhu, C.; Huang, Y.; Xi, K.; Maier, J.; Yu, Y. Challenges and Perspectives for NASICON-Type Electrode Materials for Advanced Sodium-Ion Batteries. Adv. Mater. 2017, 29, 1700431. [Google Scholar] [CrossRef]
  19. Takeda, Y.; Nakahara, K.; Nishijima, M.; Imanishi, N.; Yamamoto, O.; Takano, M.; Kanno, R. Crystal chemistry and physical properties of La2−xSrxNiO4 (0 ≤ x ≤ 1.6). Mater. Res. Bull. 1990, 25, 293–306. [Google Scholar] [CrossRef]
  20. Ma, X.; Chen, H.; Ceder, G. Electrochemical properties of monoclinic NaMnO2. J. Electrochem. Soc. 2011, 158, A1307. [Google Scholar] [CrossRef]
  21. Komaba, S.; Takei, C.; Nakayama Ogata, T.A.; Yabuuchi, N. Electrochemical intercalation activity of layered NaCrO2 vs. LiCrO2. Electrochem. Commun. 2010, 12, 355. [Google Scholar] [CrossRef]
  22. Hamani, D.; Ati, M.; Tarascon, J.M.; Rozier, P. NaxVO2 as possible electrode for Na-ion batteries. Electrochem. Commun. 2011, 13, 938. [Google Scholar] [CrossRef]
  23. Recham, N.; Chotard, J.N.; Dupont, L.; Djell, K.; Armand, M.; Tarasconz, J.M. Ionothermal synthesis of sodium-based fluorophosphate cathode materials. J. Electrochem. Soc. 2009, 156, A993. [Google Scholar] [CrossRef]
  24. Pu, X.J.; Wang, H.M.; Yuan, T.C.; Cao, S.N.; Liu, S.Y.; Xu, L.; Yang, H.X.; Ai, X.P.; Chen, Z.X.; Cao, Y.L. Na4Fe3(PO4)2P2O7/C nanospheres as low-cost, high-performance cathode material for sodium-ion batteries. Energy Stor. Mater. 2019, 22, 330–336. [Google Scholar] [CrossRef]
  25. Rajagopalan, R.; Chen, B.; Zhang, Z.; Wu, X.L.; Du, Y.; Huang, Y.; Li, B.; Zong, Y.; Wang, J.; Nam, G.H.; et al. Improved Reversibility of Fe3+/Fe4+ Redox Couple in Sodium Super Ion Conductor Type Na3Fe2(PO4)3 for Sodium-Ion Batteries. Adv. Mater. 2017, 29, 1605694. [Google Scholar] [CrossRef]
  26. Xia, X.; Cao, Y.; Yao, L.; Yang, H.; Zhang, J. MCNT-Reinforced Na3Fe2(PO4)3 as Cathode Material for Sodium-Ion Batteries. Arab J. Sci. Eng. 2020, 45, 143–151. [Google Scholar] [CrossRef]
  27. Piernas Muñoz, M.J.; Martínez, E.C. Prussian Blue Based Batteries; Springer: Cham, Switzerland, 2018; pp. 9–22. [Google Scholar]
  28. Qian, J.; Wu, C.; Cao, Y.; Ma, Z.; Huang, Y.; Ai, X.; Yang, H. Prussian blue cathode materials for sodium-ion batteries and other ion batteries. Adv. Energy Mater. 2018, 8, 1702619. [Google Scholar] [CrossRef]
  29. You, Y.; Wu, X.L.; Yin, Y.X.; Guo, Y.-G. High-quality Prussian blue crystals as superior cathode materials for room-temperature sodium-ion batteries. Energy Environ. Sci. 2014, 7, 1643. [Google Scholar] [CrossRef]
  30. Wang, B.; Han, Y.; Wang, X.; Bahlawane, N.; Pan, H.; Yan, M.; Jiang, Y. Prussian blue analogs for rechargeable batteries. Iscience 2018, 3, 110. [Google Scholar] [CrossRef] [Green Version]
  31. Fernández, N.; Sánchez-Fontecoba, P.; Castillo-Martínez, E.; Carretero-González, J.; Rojo, T.; Armand, M. Polymeric redox-active electrodes for sodium-ion batteries. Chem. Sus. Chem. 2018, 11, 311–319. [Google Scholar] [CrossRef]
  32. Mendiboure, A.; Delmas, C.; Hagenmuller, P. Electrochemical intercalation and deintercalation of NaxMnO2 bronzes. J. Solid State Chem. 1985, 57, 323–331. [Google Scholar] [CrossRef]
  33. Manikandan, P.; Heo, S.; Kim, H.W.; Jeong, H.Y.; Lee, E.; Kim, Y. Structural characterization of layered Na0.5Co0.5Mn0.5O2 material as a promising cathode for sodium-ion batteries. J. Power Source 2017, 80, 442–449. [Google Scholar] [CrossRef]
  34. Komaba, S.; Yabuuchi, N.; Nakayama, T.; Ogata, A.; Ishikawa, T.; Nakai, I. Study on the reversible electrode reaction of Na1-xNi0.5Mn0.5O2 for a rechargeable sodium-ion battery. Inorg. Chem. 2012, 51, 6211–6220. [Google Scholar] [CrossRef] [PubMed]
  35. Han, D.W.; Ku, J.H.; Kim, R.H.; Yun, D.J.; Lee, S.S.; Doo, S.G. Aluminum manganese oxides with mixed crystal structure: High-energy-density cathodes for rechargeable sodium batteries. Chem. Sus. Chem. 2014, 7, 1870–1875. [Google Scholar] [CrossRef] [PubMed]
  36. Sun, J.; Shen, J.; Wang, T. Electrochemical study of Na0.66Ni0.33Mn0.67−xMoxO2 as cathode material for sodium-ion battery. J. Alloys Compd. 2017, 709, 481–486. [Google Scholar] [CrossRef]
  37. Ding, F.X.; Zhao, C.L.; Zhou, D.; Meng, Q.S.; Xiao, D.D.; Zhang, Q.Q.; Niu, Y.S.; Li, Y.Q.; Rong, X.H.; Lu, Y.X.; et al. A Novel Ni-rich O3-Na[Ni0.60Fe0.25Mn0.15]O2 Cathode for Na-ion Batteries. Energy Storage Mater. 2020, 30, 420–430. [Google Scholar] [CrossRef]
  38. Fang, Y.J.; Yu, X.Y.; Xiong, W.L. A practical high-energy cathode for sodium-ion batteries based on uniform P2-Na0. 7CoO2 microspheres. Angew. Chem. Int. Ed. 2017, 21, 5801–5805. [Google Scholar] [CrossRef]
  39. Wang, L.; Song, J.; Qiao, R.M.; Wray, L.A.; Hossain, M.A.; Chuang, Y.D.; Yang, W.L.; Lu, Y.H.; Evans, D. Rhombohedral Prussian White as Cathode for Rechargeable Sodium-Ion Batteries. J. Am. Chem. Soc. 2015, 137, 2548–2554. [Google Scholar] [CrossRef]
  40. Hwang, J.; Kim, J. A supercritical methanol route for the synthesis of sodium iron oxide submicron plates for use as a cathode material for sodium-ion batteries. Mater. Lett. 2017, 206, 100–104. [Google Scholar] [CrossRef]
  41. Zhou, D.; Huang, W.; Lv, X.; Zhao, F. A novel P2/O3 bi phase Na0.67Fe0.425Mn0.425Mg0.15O2 as cathode for high-performance sodium-ion batteries. J. Power Source 2019, 421, 147–155. [Google Scholar] [CrossRef]
  42. Li, A.H.; Feng, Z.Y.; Sun, Y.; Shang, L.M.; Xu, L.Q. Porous organic polymer/RGO composite as high performance cathode for half and full sodium ion batteries. J. Power Source 2017, 343, 424–430. [Google Scholar] [CrossRef]
  43. Luo, C.; Fan, X.L.; Ma, Z.H.; Gao, T.; Wang, C.S. Self-Healing Chemistry between Organic Material and Binder for Stable Sodium-Ion Batteries. Chem 2017, 3, 1050–1062. [Google Scholar] [CrossRef] [Green Version]
  44. Li, J.; Chen, Y.; Zhou, T.; Shi, H.; Zheng, Z.; Wang, Y.; Guo, L. Dual-carbon coated Na3V2(PO4)3 derived from reduced graphene oxide and nanocellulose with porous structure for high performance sodium-ion batteries. Appl. Surf. Sci. 2023, 610, 155553. [Google Scholar] [CrossRef]
  45. Li, H.; Chen, X.; Jin, T.; Bao, W.; Zhang, Z.; Jiao, L. Robust graphene layer modified Na2MnP2O7 as a durable high-rate and high energy cathode for Na-ion batteries. Energy Storage Mater. 2019, 16, 383–390. [Google Scholar] [CrossRef]
  46. Zaghib, K.; Julien, C.M. Structure and electrochemistry of FePO4·2H2O hydrate. J. Power Source 2005, 142, 279–284. [Google Scholar] [CrossRef]
  47. Karami, H.; Taala, F. Synthesis, characterization and application of Li3Fe2(PO4)3 nanoparticles as cathode of lithium-ion rechargeable batteries. J. Power Source 2011, 196, 6400–6411. [Google Scholar] [CrossRef]
  48. Yin, Y.; Hu, Y.; Wu, P.; Zhang, H.; Cai, C. A graphene-amorphous FePO4 hollow nanosphere hybrid as a cathode material for lithium ion batteries. Chem. Commun. 2012, 48, 2137–2139. [Google Scholar] [CrossRef]
  49. Guo, B.; Ruan, H.; Zheng, C.; Fei, H.; Wei, M. Hierarchical LiFePO4 with a controllable growth of the (010) facet for lithiumion batteries. Sci. Rep. 2013, 3, 2788–2793. [Google Scholar] [CrossRef] [Green Version]
  50. Zhang, S.M.; Zhang, J.X.; Xu, S.J.; Yuan, X.J.; Tan, T. Synthesis, morphological analysis and electrochemical performance of iron hydroxyl phosphate as a cathode material for lithium ion batteries. J. Power Source 2013, 243, 274–279. [Google Scholar] [CrossRef]
  51. Pu, X.; Wang, H.; Zhao, D.; Yang, H.; Ai, X.; Cao, S.; Chen, Z.; Cao, Y. Recent progress in rechargeable sodium-ion batteries: Toward high-power applications. Small 2019, 15, 1805427. [Google Scholar] [CrossRef]
  52. Niu, Y.; Zhang, Y.; Xu, M. A review on pyrophosphate framework cathode materials for sodium-ion batteries. J. Mater. Chem. A 2019, 7, 15006–15025. [Google Scholar] [CrossRef]
  53. Guo, S.P.; Li, J.C.; Xu, Q.T.; Ma, Z.; Xue, H.G. Recent achievements on polyanion-type compounds for sodium-ion batteries: Syntheses, crystal chemistry and electrochemical performance. J. Power Source 2017, 361, 285–299. [Google Scholar] [CrossRef]
  54. Wang, X.; Wang, W.; Zhu, B.; Qian, F.; Fang, Z. Mo-doped Na3V2(PO4)3@C composites for high stable sodium ion battery cathode. Front. Mater. Sci. 2018, 12, 53–63. [Google Scholar] [CrossRef]
  55. Li, X.; Huang, Y.; Wang, J.; Miao, L.; Li, Y.; Liu, Y.; Qiu, Y.; Fang, C.; Han, J.; Huang, Y. High valence Mo-doped Na3V2(PO4)3/C as a high rate and stable cycle-life cathode for sodium battery. J. Mater. Chem. A 2018, 6, 1390–1396. [Google Scholar] [CrossRef]
  56. Zhu, Q.; Wang, M.; Nan, B.; Shi, H.; Zhang, X.; Deng, Y.; Wang, L.; Chen, Q.; Lu, Z. Core/shell nanostructured Na3V2(PO4)3/C/TiO2 composite nanofibers as a stable anode for sodium-ion batteries. J. Power Source 2017, 362, 147–159. [Google Scholar] [CrossRef]
  57. Zheng, W.; Huang, X.; Ren, Y.; Wang, H.; Zhou, S.; Chen, Y.; Ding, X.; Zhou, T. Porous spherical Na3V2(PO4)3/C composites synthesized via a spray drying-assisted process with high-rate performance as cathode materials for sodium-ion batteries. Solid State Ion. 2017, 308, 161–166. [Google Scholar] [CrossRef]
  58. Li, G.; Jiang, D.; Wang, H.; Lan, X.; Zhong, H.; Jiang, Y. Glucose-assisted synthesis of Na3V2(PO4)3 composite as an electrode material for high-performance sodium-ion batteries. J. Power Source 2014, 265, 325–334. [Google Scholar] [CrossRef]
  59. Tao, S.; Cui, P.X.; Huang, W.F.; Yu, Z.; Wang, X.B.; Wei, S.H.; Liu, D.B.; Song, L.; Chu, W.S. Sol-gel design strategy for embedded Na3V2(PO4)3 particles into carbon matrices for high-performance sodium-ion batteries. Carbon 2016, 96, 1028–1033. [Google Scholar] [CrossRef]
  60. Zhu, C.B.; Song, K.P.; Aken, P.A.; Maier, V.J.; Yu, Y. Carbon-coated Na3V2(PO4)3 embedded in porous carbon matrix: An ultrafast Na-storage cathode with the potential of outperforming Li cathodes. Nano Lett. 2014, 14, 2175–2180. [Google Scholar] [CrossRef]
  61. Li, S.; Dong, Y.F.; Xu, L.; Xu, X.; He, L.; Mai, L.Q. Effect of carbon matrix dimensions on the electrochemical properties of Na3V2(PO4)3 nanograins for high-performance symmetric sodium-ion batteries. Adv. Mater. 2014, 26, 3358–3363. [Google Scholar] [CrossRef]
  62. Zhu, X.M.; Fang, Y.J.; Ai, X.P.; Yang, H.X.; Cao, Y.L. Na3V2(PO4)3/C nanocomposite synthesize via pre-reduction process as high-performance cathode material for sodium-ion batteries. J. Alloys Compd. 2015, 646, 170–174. [Google Scholar] [CrossRef]
  63. Jian, Z.L.; Zhao, L.; Pan, H.L.; Hu, Y.S.; Li, H.; Chen, W.; Chen, L.Q. Carbon coated Na3V2(PO4)3 as novel electrode material for sodium ion batteries. Electrochem. Commun. 2012, 14, 86–89. [Google Scholar] [CrossRef]
  64. Cheng, D.; Li, Y.; Zhang, J.; Tian, M.; Wang, B.; He, Z.; Dai, L.; Wang, L. Recent advances in electrospun carbon fiber electrode for vanadium redox flow battery: Properties, structures, and perspectives. Carbon 2020, 170, 527–542. [Google Scholar] [CrossRef]
  65. Jiang, Y.; Zeng, L.; Wang, J.; Li, W.; Pana, F.; Yu, Y. A carbon coated NASICON structure material embedded in porous carbon enabling superior sodium storage performance: NaTi2(PO4)3 as an example. Nanoscale 2015, 7, 14723. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, Q.; Wang, W.; Wang, Y.; Feng, P.; Wang, K.; Cheng, S.; Jiang, K. Controllable construction of 3D-skeleton-carbon coated Na3V2(PO4)3 for high-performance sodium ion battery cathode. Nano Energy 2016, 20, 11. [Google Scholar] [CrossRef]
  67. Fang, J.; Wang, S.; Li, Z.; Chen, H.; Xia, L.; Ding, L.; Wang, H. Porous Na3V2 (PO4)3@C nanoparticles enwrapped in three-dimensional graphene for high performance sodium-ion batteries. J. Mater. Chem. A 2016, 4, 1180. [Google Scholar] [CrossRef]
  68. Li, H.; Yu, X.; Bai, Y.; Wu, F.; Wu, C.; Liu, L.Y.; Yang, X.Q. Effects of Mg doping on the remarkably enhanced electrochemical performance of Na3V2(PO4)3 cathode materials for sodium ion batteries. J. Mater. Chem. A 2015, 3, 9578. [Google Scholar] [CrossRef]
  69. Aragón, M.J.; Lavela, P.; Alcántara, R.; Tirado, J.L. Effect of aluminum doping on carbon loaded Na3V2(PO4)3 as cathode material for sodium-ion batteries. Electrochim. Acta 2015, 180, 824. [Google Scholar] [CrossRef]
  70. Zhu, C.; Kopold, P.; van Aken, P.A.; Maier, J.; Yu, Y. High power-high energy sodium battery based on threefold interpenetrating network. Adv. Mater. 2016, 28, 2409. [Google Scholar] [CrossRef]
  71. Song, W.; Ji, X.; Wu, Z.; Zhu, Y.; Yang, Y.; Chen, J.; Jing, M.; Li, F.; Banks, C.E. First exploration of Na-ion migration pathways in the NASICON structure Na3V2(PO4)3. J. Mater. Chem. A 2014, 2, 5358. [Google Scholar] [CrossRef]
  72. Xiao, H.; Huang, X.; Ren, Y.; Wang, H.; Ding, J.; Zhou, S.; Ding, X.; Chen, Y. Enhanced sodium ion storage performance of Na3V2(PO4)3 with N-doped carbon by folic acid as carbon-nitrogen source. J. Alloys Compd. 2018, 732, 454–459. [Google Scholar] [CrossRef]
  73. Klee, R.; Aragon, M.J.; Lavela, P.; Alcantara, R.; Tirado, J.L. Na3V2(PO4)3/C nanorods with improved electrode-electrolyte interface as cathode material for sodium-ion batteries. ACS Appl. Mater. Interfaces 2016, 8, 23151–23159. [Google Scholar] [CrossRef] [PubMed]
  74. Lim, S.Y.; Kim, H.; Shakoor, R.A.; Jung, Y.; Choi, J.W. Electrochemical and Thermal Properties of NASICON Structured Na3V2(PO4)3 as a Sodium Rechargeable Battery Cathode: A Combined Experimental and Theoretical Study. J. Electrochem. Soc. 2012, 159, A1393–A1397. [Google Scholar] [CrossRef]
  75. Bianchini, M.; Fauth, F.; Brisset, N.; Weill, F.; Masquelier, E.; Croguennec, L. Comprehensive investigation of the Na3V2(PO4)2F3–NaV2(PO4)2F3 system by operando high resolution synchrotron X-ray diffraction. Chem. Mater. 2015, 27, 3009–3020. [Google Scholar] [CrossRef]
  76. Liu, J.; Tang, K.; Song, K.; van Aken, P.A.; Yu, Y.; Maier, J. Electrospun Na3V2(PO4)3/C nanofibers as stable cathode materials for sodium-ion batteries. Nanoscale 2014, 6, 5081–5086. [Google Scholar] [CrossRef]
  77. Xiao, H.; Huang, X.; Ren, Y.; Ding, X.; Chen, Y.; Zhou, S. Improved electrochemical performance of Na3V2(PO4)3 modified by N-doped carbon coating and Zr doping. Solid State Ion. 2019, 333, 93–100. [Google Scholar] [CrossRef]
  78. 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. [Google Scholar] [CrossRef]
  79. Chang, X.Q.; Zhu, Q.Z.; Sun, N.; Guan, Y.B.; Wang, R.; Zhao, J.H.; Feng, M.Y.; Xu, B. Graphene-bound Na3V2(PO4)3 film electrode with excellent cycle and rate performance for Na-ion batteries. Electrochim. Acta 2018, 296, 282–290. [Google Scholar] [CrossRef]
  80. Xu, J.Y.; Gu, E.L.; Zhang, Z.Z.; Xu, Z.H.; Xu, Y.F.; Du, Y.C.; Zhu, X.S.; Zhou, X.S. Fabrication of porous Na3V2(PO4)3/reduced graphene oxide hollow spheres with enhanced sodium storage performance. J. Colloid Interface Sci. 2020, 567, 84–91. [Google Scholar] [CrossRef]
  81. Xu, Y.N.; Wei, Q.L.; Xu, C.; Li, Q.D.; An, Q.Y.; Zhang, P.F.; Sheng, J.Z.; Zhou, L.; Mai, L.Q. Layer-by-Layer Na3V2(PO4)3 Embedded in Reduced Graphene Oxide as Superior Rate and Ultralong-Life Sodium-Ion Battery Cathode. Adv. Energy Mater. 2016, 6, 1600389. [Google Scholar] [CrossRef]
  82. Shen, W.; Li, H.; Guo, Z.; Wang, C.; Li, Z.; Xu, Q.; Liu, H.; Wang, Y.; Xia, Y. Double-Nanocarbon Synergistically Modified Na3V2(PO4)3: An Advanced Cathode for High-Rate and Long-Life Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 15341–15351. [Google Scholar] [CrossRef] [PubMed]
  83. Chen, H.; Zhang, B.; Wang, X.; Dong, P.; Tong, H.; Zheng, J.C.; Yu, W.; Zhang, J. CNT-Decorated Na3V2(PO4)3 Microspheres as a High-Rate and Cycle-Stable Cathode Material for Sodium Ion Batteries. ACS Appl. Mater. Interfaces 2018, 10, 3590–3595. [Google Scholar] [CrossRef] [PubMed]
  84. Sun, S.Q.; Chen, Y.J.; Cheng, J.; Tian, Z.Y.; Wang, C.; Wu, G.P.; Liu, C.C.; Wang, Y.Z.; Li, G. Constructing dimensional gradient structure of Na3V2(PO4)3/C@CNTs-WC by wolfram substitution for superior sodium storage. Chem. Eng. J. 2021, 420, 130453. [Google Scholar] [CrossRef]
  85. Duan, W.C.; Zhu, Z.Q.; Li, H.; Hu, Z.; Zhang, K.; Cheng, F.Y.; Chen, J. Na3V2(PO4)3@C core–shell nanocomposites for rechargeable sodium-ion batteries. J. Mater. Chem. A 2014, 2, 8668. [Google Scholar] [CrossRef]
  86. Li, S.J.; Ge, P.; Zhang, C.T.; Sun, W.; Hou, H.S.; Ji, X.B. The electrochemical exploration of double carbon-wrapped Na3V2(PO4)3: Towards long-time cycling and superior rate sodium-ion battery cathode. J. Power Source 2017, 366, 249–258. [Google Scholar] [CrossRef]
  87. Jiang, Y.; Yang, Z.; Li, W.; Zeng, L.; Pan, F.; Wang, M.; Wei, X.; Hu, G.; Gu, L.; Yu, Y. Nanoconfined Carbon-Coated Na3V2(PO4)3 Particles in Mesoporous Carbon Enabling Ultralong Cycle Life for Sodium-Ion Batteries. Adv. Energy Mater. 2015, 5, 1402104. [Google Scholar] [CrossRef]
  88. Wang, G.; Liu, H.; Liu, J.; Qiao, S.; Lu, G.M.; Munroe, P.; Ahn, H. Mesoporous LiFePO4/C Nanocomposite Cathode Materials for High Power Lithium Ion Batteries with Superior Performance. Adv. Mater. 2010, 22, 4944. [Google Scholar] [CrossRef]
  89. Zhang, H.; Tao, H.; Jiang, Y.; Jiao, Z.; Wu, M.; Zhao, B. Ordered CoO/CMK-3 nanocomposites as the anode materials for lithium-ion batteries. J. Power Source 2010, 195, 2950–2955. [Google Scholar] [CrossRef]
  90. Hong, X.D.; Huang, X.B.; Ren, Y.R.; Wang, H.Y.; Ding, X.; Jin, J.L. Superior Na-storage performance of Na3V2(PO4)3/C-Ag composites as cathode material for Na-ion battery. J. Alloys Compd. 2020, 822, 153587. [Google Scholar] [CrossRef]
  91. Wang, Y.P.; Zhao, K.J.; Wang, k.; Li, H.H.; Jiang, H.B.; Chen, L. N-doped carbon confined Na3V2(PO4)3 derived from an organophosphonic acid as a high-performance cathode for sodium-ion batteries. J. Alloys Compd. 2020, 844, 156118. [Google Scholar] [CrossRef]
  92. Jiang, Y.; Wu, Y.; Chen, Y.; Qi, Z.; Shi, J.; Gu, L.; Yu, Y. Design Nitrogen (N) and Sulfur (S) co-Doped 3D Graphene Network Architectures for High-Performance Sodium Storage. Small 2018, 14, 1703471. [Google Scholar] [CrossRef] [PubMed]
  93. Ren, W.H.; Yao, X.H.; Niu, C.J.; Zheng, Z.P.; Zhao, K.N.; An, Q.Y.; Wei, Q.L.; Yan, M.Y.; Zhang, L.; Mai, L.Q. Cathodic polarization suppressed sodium-ion full cell with a 3.3 V high-voltage. Nano Energy 2016, 28, 216–223. [Google Scholar] [CrossRef]
  94. Kajiyama, S.; Kikkawa, J.; Hoshino, J.; Okubo, M.; Hosono, E. Assembly of Na3V2(PO4)3 Nanoparticles Confined in a One-Dimensional Carbon Sheath for Enhanced Sodium-Ion Cathode Properties. Chem. Eur. J. 2014, 20, 12636–12640. [Google Scholar] [CrossRef] [PubMed]
  95. Shen, W.; Wang, C.; Xu, Q.J.; Liu, H.M.; Wang, Y.G. Nitrogen-Doping-Induced Defects of a Carbon Coating Layer Facilitate Na-Storage in Electrode Materials. Adv. Energy Mater. 2015, 5, 1400982. [Google Scholar] [CrossRef]
  96. Yi, H.M.; Li, D.; Lv, Z.Q.; Li, R.; Ling, M.; Zhang, H.; Zheng, Q.; Li, X. Constructing high-performance 3D porous self-standing electrodes with various morphologies and shapes by a flexible phase separation-derived method. J. Mater. Chem. A 2019, 7, 22550–22558. [Google Scholar] [CrossRef]
  97. Shen, W.; Li, H.; Wang, C.; Li, Z.H.; Xu, Q.J.; Liu, H.M.; Wang, Y.G. Improved electrochemical performance of the Na3V2(PO4)3 cathode by B-doping of the carbon coating layer for sodium-ion batteries. J. Mater. Chem. A 2015, 3, 15190–15201. [Google Scholar] [CrossRef]
Figure 1. The crystal structure of Na3V2(PO4)3.
Figure 1. The crystal structure of Na3V2(PO4)3.
Electronicmat 04 00003 g001
Figure 2. (a) Schematic illustration of 3D hierarchical meso- and macroporous NVP@C@rGO cathode with pathways for both electrons and sodium ions, (b) schematic illustration of preparing NVP@C@rGO composite, the (c) rate performance, (d) discharge profiles at various current rates and (e) EIS spectra of NVP@C@rGO. Reprinted with permission from Ref. [78]. 2015, Wiley−VCH. The (f) SEM image graphene-bound NVP, (g) rate performance, and (h) EIS spectra of PVDF-bound NVP and graphene-bound NVP. Reprinted with permission from Ref. [79]. 2015, Elsevier.
Figure 2. (a) Schematic illustration of 3D hierarchical meso- and macroporous NVP@C@rGO cathode with pathways for both electrons and sodium ions, (b) schematic illustration of preparing NVP@C@rGO composite, the (c) rate performance, (d) discharge profiles at various current rates and (e) EIS spectra of NVP@C@rGO. Reprinted with permission from Ref. [78]. 2015, Wiley−VCH. The (f) SEM image graphene-bound NVP, (g) rate performance, and (h) EIS spectra of PVDF-bound NVP and graphene-bound NVP. Reprinted with permission from Ref. [79]. 2015, Elsevier.
Electronicmat 04 00003 g002
Figure 3. (a) Schematic illustration for the fabrication of NVP/rGO HSs, the (b) SEM, and (c) TEM images of NVP/rGO HSs. Reprinted with permission from Ref. [80]. 2020, Elsevier.
Figure 3. (a) Schematic illustration for the fabrication of NVP/rGO HSs, the (b) SEM, and (c) TEM images of NVP/rGO HSs. Reprinted with permission from Ref. [80]. 2020, Elsevier.
Electronicmat 04 00003 g003
Figure 4. The (a) structure, TEM images (b,c), and (d) rate performance of NVP@C+N@CNTs. Reprinted with permission from Ref. [82]. 2016, American Chemical Society.
Figure 4. The (a) structure, TEM images (b,c), and (d) rate performance of NVP@C+N@CNTs. Reprinted with permission from Ref. [82]. 2016, American Chemical Society.
Electronicmat 04 00003 g004
Figure 5. The SEM images of (a) NVP/C, (b) NVP/C5, (c) NVP/C10, and (d) NVP/C15, (e) TEM and (f) HRTEM images of NVP/C10, (g) charge/discharge curves of different electrodes at 110 mA/g, the inset is the rate performance, and (h) EIS spectra of NVP/C and NVP/C10. Reprinted with permission from Ref. [83]. 2018, American Chemical Society.
Figure 5. The SEM images of (a) NVP/C, (b) NVP/C5, (c) NVP/C10, and (d) NVP/C15, (e) TEM and (f) HRTEM images of NVP/C10, (g) charge/discharge curves of different electrodes at 110 mA/g, the inset is the rate performance, and (h) EIS spectra of NVP/C and NVP/C10. Reprinted with permission from Ref. [83]. 2018, American Chemical Society.
Electronicmat 04 00003 g005
Figure 6. The (a) and (b) TEM images, (c) cycling performance, (d) EIS spectra, and (e) long-term cycling performance of NVPC-800. Reprinted with permission from Ref. [86]. 2019 Elsevier B.V.
Figure 6. The (a) and (b) TEM images, (c) cycling performance, (d) EIS spectra, and (e) long-term cycling performance of NVPC-800. Reprinted with permission from Ref. [86]. 2019 Elsevier B.V.
Electronicmat 04 00003 g006
Figure 7. (a) Schematical illustrations of fabricating NVP@C@CMK-3 cathode and the electrochemical reactions confined within a single carbon pore, the (b) charge/discharge curves and (c) EIS spectra of NVP@C@CMK-3. Reprinted with permission from Ref. [87]. 2015, Wiley-VCH.
Figure 7. (a) Schematical illustrations of fabricating NVP@C@CMK-3 cathode and the electrochemical reactions confined within a single carbon pore, the (b) charge/discharge curves and (c) EIS spectra of NVP@C@CMK-3. Reprinted with permission from Ref. [87]. 2015, Wiley-VCH.
Electronicmat 04 00003 g007
Figure 8. (a) Schematic illustration of the formation of NVP/C-Ag, (b) XRD patterns, (c) TG curves, and (d) charge/discharge curves at various current densities of NVP/C-Ag. Reprinted with permission from Ref. [90]. 2020, Elsevier B.V. (e) Schematical synthesis of NVP-NC, (f) the SEM image of NVP-NC Reprinted with permission from Ref. [91]. 2020, Elsevier B.V.
Figure 8. (a) Schematic illustration of the formation of NVP/C-Ag, (b) XRD patterns, (c) TG curves, and (d) charge/discharge curves at various current densities of NVP/C-Ag. Reprinted with permission from Ref. [90]. 2020, Elsevier B.V. (e) Schematical synthesis of NVP-NC, (f) the SEM image of NVP-NC Reprinted with permission from Ref. [91]. 2020, Elsevier B.V.
Electronicmat 04 00003 g008
Figure 9. (a) Schematical synthesis of NVP@C@3DPGFs-NS, the (b) SEM image, (c) TEM image, (d) charge/discharge curves, and (e) cycling capability of NVP@C@3DPGFs-NS. Reprinted with permission from Ref. [92]. 2017, WILEY-VCH.
Figure 9. (a) Schematical synthesis of NVP@C@3DPGFs-NS, the (b) SEM image, (c) TEM image, (d) charge/discharge curves, and (e) cycling capability of NVP@C@3DPGFs-NS. Reprinted with permission from Ref. [92]. 2017, WILEY-VCH.
Electronicmat 04 00003 g009
Table 1. The comparison of sodium and lithium.
Table 1. The comparison of sodium and lithium.
MetalNaLi
Atoms radius/Å1.861.52
Ionic radius/Å1.020.76
Standard electrode potential/V−2.71−3.04
Earth crust abundance2.8%0.0065%
Theoretical specific capacity/(mah/g)11663862
Table 2. The comparison of performance among different cathodes for SIBs.
Table 2. The comparison of performance among different cathodes for SIBs.
MaterialsVoltageReversible CapacityRate PerformanceRef.
NaMnO22.59 VC/30, 149 mAh/gC/10, 144 mAh/g[32]
Na0.5Co0.5Mn0.5O22.8 VC/2, 141 mAh/g5 C, 80 mAh/g[33]
NaNi0.5Mn0.5O2>2.5 VC/50, 125 mAh/gC/30, 100 mAh/g[34]
NaAl0.1Mn0.9O22.56 V-C/10, 145 mAh/g[35]
Na0.66Ni0.33Mn0.62Mo0.05O23.25 V34 mA/g, 125 mAh/g1 C, 112 mAh/g[36]
NaNi0.60Fe0.25Mn0.15O23.1 VC/2, 190 mAh/g2 C, 134 mAh/g[37]
Na0.7CoO22.8 V0.04 C, 125 mAh/g16 C, 64 mAh/g[38]
Na1.92Fe[Fe(CN)6]3.2 VC/10, 157 mAh/g10 C, 145 mAh/g[39]
NaFeO23.4 V10 mA/g, 94 mAh/g-[40]
Na0.67Fe0.425Mn0.425Mg0.15O23.6 VC/10, 146.6 mAh/g2 C, 58.4 mAh/g[41]
Na2PDHBQS/RGO1.4 VC/10, 228 mAh/g4 C, 147 mAh/g[42]
Na2C6O62.1 VC/10, 190 mAh/g10 C, 95 mAh/g[43]
CNF-GC-NVP3.4 V1 C, 123.77 mAh/g60 C, 95.59 mAh/g[44]
Na2MnP2O73.6 VC/10, 94 mAh/g10 C, 55 mAh/g[45]
Table 3. Summary of NVP with different morphology for storage properties.
Table 3. Summary of NVP with different morphology for storage properties.
CompositionMorphologyRate PerformanceCycle NumberRef.
Na3V2(PO4)3@ACNanoparticle5 C, 100.6 mAh/g200[44]
C@Na3V2(PO4)3@pCDouble-shell nanospheres10 C, 103 mAh/g1000[60]
Na3V2(PO4)3-F/CNanofibersC/10, 103 mAh/g50[76]
NVP@C@rGO3D porous composites1 C, 115 mAh/g10,000[78]
Graphene-bound NVP3D continuous network30 C, 71.9 mAh/g1000[79]
NVP/rGO HSsHollow spheres1 C, 109 mAh/g400[80]
NVP@C+N@CNTsNanoparticle with CNTsC/5, 86.5 mAh/g400[82]
NVP/CNTMicrosphere10 C, 88.5 mAh/g150[83]
NVP/C@CNTs-WCNanoparticle with CNTs50 C, 84.5 mAh/g400[84]
NVP@CDouble carbon-wrapped1 C, 100 mAh/g40[85]
Na3V2(PO4)3@C@CMK-33D CMK-31 C, 115 mAh/g2000[87]
NVP/C-AgParticle10 C, 95 mAh/g500[90]
NVP-NCParticle10 C, 94 mAh/g600[91]
Na3V2(PO4)3@C@3DPGFs-NS3D porous composites20 C, 86 mAh/g12,000[92]
Na3V2(PO4)3@C3D nanofibers10 C, 110 mAh/g1000[93]
Na3V2(PO4)3/CSCore-sheath nanowires1 C, 94 mAh/g50[94]
Na3V2(PO4)3@CNT-G3D continuous free-standing foam30 C, 109 mAh/g2000[95]
Na3V2(PO4)3@C3D porous self-standingC/2,110 mAh/g2000[96]
Na3V2(PO4)3@C-B0.38NanoparticlesC/5, 87.1 mAh/g50[97]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

He, Y.; Li, H. Recent Research Process of Carbon Engineering on Na3V2(PO4)3 for Sodium-Ion Battery Cathodes: A Mini Review. Electron. Mater. 2023, 4, 17-32. https://doi.org/10.3390/electronicmat4010003

AMA Style

He Y, Li H. Recent Research Process of Carbon Engineering on Na3V2(PO4)3 for Sodium-Ion Battery Cathodes: A Mini Review. Electronic Materials. 2023; 4(1):17-32. https://doi.org/10.3390/electronicmat4010003

Chicago/Turabian Style

He, Yaxuan, and Haibo Li. 2023. "Recent Research Process of Carbon Engineering on Na3V2(PO4)3 for Sodium-Ion Battery Cathodes: A Mini Review" Electronic Materials 4, no. 1: 17-32. https://doi.org/10.3390/electronicmat4010003

Article Metrics

Back to TopTop