Next Article in Journal
A Comparative Study of the Band-Edge Exciton Fine Structure in Zinc Blende and Wurtzite CdSe Nanocrystals
Next Article in Special Issue
Rational Construction of C@Sn/NSGr Composites as Enhanced Performance Anodes for Lithium Ion Batteries
Previous Article in Journal
Terahertz Emission Spectroscopy of Ultrafast Coupled Spin and Charge Dynamics in Nanometer Ferromagnetic Heterostructures
Previous Article in Special Issue
Oxygen Vacancies in Bismuth Tantalum Oxide to Anchor Polysulfide and Accelerate the Sulfur Evolution Reaction in Lithium–Sulfur Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vanadium Hexacyanoferrate as a High-Capacity and High-Voltage Cathode for Aqueous Rechargeable Zinc Ion Batteries

1
School of Science, Dalian Maritime University, Dalian 116026, China
2
School of Physics and Materials Engineering, Dalian Minzu University, Dalian 116600, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(23), 4268; https://doi.org/10.3390/nano12234268
Submission received: 3 November 2022 / Revised: 28 November 2022 / Accepted: 29 November 2022 / Published: 30 November 2022

Abstract

:
Prussian blue analogs (PBAs) are widely used as electrode materials for secondary batteries because of their cheapness, ease of synthesis, and unique structural properties. Nevertheless, the unsatisfactory capacity and cyclic stability of PBAs are seriously preventing their practical applications. Here, vanadium hexacyanoferrate (VHCF) is successfully prepared and used as a cathode for aqueous zinc-ion batteries (AZIBs). When using 3 M Zn(CF3SO3)2 as the electrolyte, a high capacity of ~230 mA h g−1 and a high voltage of ~1.2 V can be achieved. The XRD result and XPS analysis indicate that the outstanding Zn2+ storage capability is due to the presence of dual electrochemical redox centers in VHCF (Fe2+ ⇋ Fe3+ and V5+ ⇋ V4+ ⇋ V3+). However, the battery shows a short cycle life (7.1% remaining capacity after 1000 cycles) due to the dissolution of VHCF. To elongate the cycle life of the battery, a high-concentration hybrid electrolyte is used to reduce the activity of water molecules. The improved battery exhibits an impressive capacity of 235.8 mA h g−1 and good capacity retention (92.9%) after 1000 cycles.

Graphical Abstract

1. Introduction

In recent years, there have been growing calls from various industries to improve battery safety, mainly due to concerns about the fire and even explosion risks of lithium-ion batteries [1,2,3]. Therefore, it is very important to develop alternative high-performance batteries with lower cost and naturally non-explosive properties [4]. Rechargeable aqueous zinc-ion batteries (AZIBs) have recently attracted much attention due to their numerous advantages, especially their excellent safety [5,6,7]. As the anode electrode of AZIBs, zinc metal has many unique merits, such as wide distribution of resources, high stability, high gravimetric capacity, and appropriate electrochemical potential [8,9]. However, the capacity and electrochemical reversibility of the cathode materials are greatly limited due to the poor electrochemical kinetics of Zn2+ [10]. Moreover, the practical application of AZIBs is still challenging due to the zinc dendrite growth and loss of active materials in conventional ZnSO4 and Zn(CF3SO3)2 aqueous electrolytes [11].
Currently, the properties of manganese-based and vanadium-based cathodes for AZIBs have been widely studied [12,13]. However, the dissolution of Mn2+ from manganese-based cathodes during the charge and discharge usually leads to a fast capacity decay [14]. Although vanadium-based cathodes have a relatively long life cycle, they generally show low operational voltage (<0.9 V), resulting in a low energy density of full batteries [15]. As another promising cathode material for AZIBs, Prussian blue analogs (PBAs) are inexpensive, high-performance, and easy to implement in industrial large-scale synthesis [16,17,18,19]. The general formula of PBAs is AxM1[M2(CN)6]y·zH2O, where the A is the alkaline cations and the M1/M2 are the transition metal ions. The most typical atom on the M2 site is Fe. The FeC6 octahedrons and the M1N4 tetrahedrons are connected through CN ligands, forming a large three-dimensional framework, enabling rapid and highly reversible intercalation/de-intercalation of cations (NH4+, Li+, Mg2+, Zn2+, etc.) [20,21]. Moreover, the electrochemical behaviors of PBAs, such as their capacity and operational voltage, can be regulated by controlling their chemical compositions. For instance, Dr. Zhang studied the properties of a zinc//zinc hexacyanoferrate (Zn//ZnHCF) battery in an aqueous electrolyte [22]. The battery delivered a capacity of 65.4 mA h g−1 at 1C with a high operational voltage of 1.7 V, and a capacity retention of 76% after 100 cycles at 1C. Later, copper hexacyanoferrate (CuHCF) was studied as a cathode electrode in AZIBs, which showed a lower operational voltage (~1.6 V) and a similar initial capacity (~64 mA h g−1) [23]. Recently, a reduced graphene oxide (RGO) modified NiHCF/RGO composite cathode for AZIBs was proposed by Xue and co-workers, giving a good cyclic stability of 1000 cycles (80.3% capacity retention). However, it only showed a limited capacity (47.8 mA h g−1) and a low discharge voltage (~1.2 V) [24]. It can be found that, as a cathode electrode for AZIBs, PBAs usually exhibit relatively low specific capacity. This is mainly due to the single electrochemical active site (Fe ions in Fe(CN)6 groups) in these PBAs. A major breakthrough in achieving high-capacity Zn//PBA batteries was achieved by Prof. Zhi and co-workers in 2019. Their proposed cobalt hexacyanoferrate (CoHCF) cathode could realize a two-species redox reaction (Co and Fe), providing a high average voltage of 1.7 V and an impressive capacity of ~170 mA h g−1 [25]. Therefore, regulating the composition and electrochemically active species of PBAs may be a feasible method to enhance the performances of Zn//PBA batteries.
Here, vanadium hexacyanoferrate (VHCF) was successfully synthesized by a co-precipitation reaction and applied to AZIBs as a cathode. The VHCF cathode showed a high capacity (~230 mA h g−1) after a short activation process. The XPS results indicate that the high capacity of VHCF was due to the existence of dual electrochemical redox centers (Fe2+ ⇋ Fe3+ and V5+ ⇋ V4+ ⇋ V3+), which can realize a three-electron redox reaction. However, its capacity decayed rapidly as the active material dissolved into the electrolyte (3 M Zn(CF3SO3)2) during cycling. Then, a high-concentration hybrid electrolyte (3 M Zn(CF3SO3)2 + 10 M LiTFSI) was used to improve the cycle stability of the Zn//VHCF battery. This improved battery exhibited a high capacity of 235.8 mA h g−1 and a remarkable capacity retention (92.9%) after 1000 cycles.

2. Materials and Methods

2.1. Synthesis of VHCF

Firstly, 25 mL of 32% hydrochloric acid was diluted to 37.5 mL with deionized (DI) water. Then, 2 g of V2O5 powders were added into the diluted acid solution with constant stirring to obtain a yellow suspension. After stirring for 1 h, 350 μL of glycerol was added drop by drop under stirring until a clear solution was formed. Then, 4.7 mL of the clear solution was diluted to 25 mL. After this, 0.5926 g K3Fe(CN)6 was dissolved in 25 mL of DI water. Then, it was slowly added to the diluted solution. After stirring for 24 h, the precipitate was centrifuged and washed six times alternately with DI water and alcohol. Finally, the prepared VHCF sample was dried in the oven at 60 °C for one day.

2.2. Materials Characterizations

The structures of the VHCF powder and cathode electrodes were studied by powder X-ray diffraction (XRD) tests using an X-ray diffractometer (D/MAX-Ultima, Shimadzu Corporation, Kyoto, Japan) with Co Kα radiation at room temperature. The valence states of the elements in the VHCF cathodes were studied by X-ray photoelectron spectroscopy (XPS) spectra using an X-ray photoelectron spectrometer (ESCALAB 250Xi, Shimadzu Corporation, Kyoto, Japan). The microstructure of the VHCF powder was observed by means of scanning electron microscopy (SEM) using a field-emission microscope (SUPRA 55 SAPPHIRE, Carl Zeiss Corporation, Jena, Germany). The element distribution in the VHCF sample was detected by the above field-emission microscope equipped with an energy dispersive spectrometer (EDS). The microstructure of the sample was checked by means of transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) measurements using an electron microscope (JEOL: 2100F, Joel Corporation, Tokyo, Japan). The thermal gravimetric (TG) curve was conducted using a thermal gravimetric analyzer (Mettler-Toledo, Zurich, Switzerland) in the air at a heating speed of 10 °C min−1.

2.3. Electrochemical Experiments

The electrochemical performance was tested using the LAND CT2001 battery tester. The voltage window of the electrochemical tests was 0.2–2.0 V. Coin cells (2032) were used to test the electrochemical properties of VHCF. A piece of Zn metal foil (100 μm thick) polished with sandpaper (2000 mesh) was used as the anode. The cathode was composed of VHCF powder at 70 wt%, Super P at 20 wt%, and a polyvinylidene fluoride (PVDF) binder at 10 wt%. To prepare the cathode electrodes, a slurry of VHCF, Super P, and PVDF was ground in N methyl-2-pyrrolidone (NMP) and then pasted onto a stainless-steel current collector. The electrode was dried in an oven at 60 °C for one day and cut into a circle using a mechanical punching machine. The diameter of the electrode is 8 mm. The loading mass of VHCF on the cathode was ~1 mg cm−2. A glass fiber filter (Whatman GF/C) was used as the separator. The electrolytes were 3 M Zn(CF3SO3)2 and 3 M Zn(CF3SO3)2 + 10 M LiTFSI aqueous solutions.

3. Results and Discussion

As shown in Figure 1a, all of the identified XRD peaks of the sample were indexed to the cubic phase of V1.5Fe(CN)6 (JCPDS No. 42-1440). Although the accurate crystal structure of VHCF is not well-studied in the literature so far, it can be expected that VHCF may exhibit similar properties to PBAs. In the VHCF crystallographic structure, V and Fe are connected by a shared CN ligand [26]. The formed porous 3D framework structure and large interstitial space can facilitate the reversible intercalation/de-intercalation of foreign cations. As shown in Figure 1b, VHCF consists of many irregular nanoparticles. The TEM image (Figure 1c) further shows that these nanoparticles are between 20 and 50 nm in size. Figure 1d is the TG curve of VHCF. The weight loss of the sample can be divided into two stages: the first stage below 200 °C is due to the evaporation of interstitial water (25.2%), and the second stage from 200 °C to 400 °C is ascribed to the composite decomposition of VHCF (21.7%) [26,27]. The EDS images (Figure 1e) confirm the even distribution of C, N, O, K, V, and Fe elements in VHCF. According to the inductively coupled plasma optical emission spectroscopy (ICP-OES) result, the molar ratio of K/V/Fe is 4.6/1.3/1.0, indicating that the possible chemical formula of VHCF is K14V4O9[Fe(CN)6]3·21.4H2O. In the VHCF crystallographic structure, 1/6 of the C=N groups are replaced by O atoms due to the presence of V=O ions, and part of the Fe(CN)6 complex sites are vacant because of the non-stoichiometry of V and Fe ions. This results in the formation of a large number of vacancies in VHCF, which are occupied by oxygen-terminating bonding. The massive void spaces and vacancies in the VHCF structure can accommodate large amounts of K ions and interstitial water molecules.
The electrochemical properties of VHCF were first studied using 3 M Zn(CF3SO3)2 as the electrolyte. Figure 2a presents the initial three charge/discharge profiles of the VHCF cathode at 0.2 A g−1. The first discharge capacity is 146.2 mA h g−1, and it gradually increases to 174.0 mA h g−1 in the third cycle due to the activation process [28,29]. The operational voltage is about 1.2 V, which is much higher than that of other V-based cathode materials. The coulombic efficiencies of the first three cycles are 87.9%, 83.1%, and 90.1%, respectively. The low coulombic efficiency indicates that the battery may experience a loss of active material during cycling. As depicted in Figure 2b, the cycling performance of VHCF was evaluated at 0.2 A g−1. The discharge capacity gradually increases from 146.2 mA h g−1 to 230.3 mA h g−1 in 15 cycles, which is the highest capacity compared with other Zn//PBA batteries reported in the literature, indicating that the introduction of vanadium into the framework can greatly improve the redox activity of PBAs. However, the capacity retention after 100 cycles is only 56.7% (compared to the highest capacity reached during cycling). The fast capacity decay is most likely caused by the loss of VHCF during cycling. A rate capability test was performed to exhibit the capacities of the Zn//VHCF battery at higher current densities. As shown in Figure 2c, 221.5, 193.3, 163.3, 149.7, 142.9, and 133.6 mA h g−1 are obtained at 0.2, 0.5, 1.0, 2.0, 3.0, and 5.0 A g−1, respectively. It can be found that a high remaining capacity of 60.3% is achieved even with a 25-fold current density increase. Such a remarkable rate performance is mainly owing to the unique framework structure and intrinsic fast reaction kinetics of VHCF. Then, the long-cycling stability of VHCF was investigated at 5 A g−1 (Figure 2d, the first three cycles were performed at 0.2 A g−1). The cell shows a negligible capacity of 10.1 mA h g−1 after 1000 cycles, resulting in a low capacity retention of 7.1%. In addition, its coulombic efficiency gradually decreases to 80% after 1000 cycles. These results suggest that VHCF is a promising cathode material for AZIBs due to its excellent Zn2+ storage capability. However, the bad cyclic stability of VHCF in the conventional 3 M Zn(CF3SO3)2 electrolyte limits its practical application. The cycling performance of the Zn//VHCF battery was also tested with a narrow voltage window between 1.2 V and 2.0 V. As shown in Figure S1, the battery shows good cycling stability. This indicates that the limited cycling life of the Zn//VHCF battery with a low cutoff voltage of 0.2 V is probably due to the dissolution of the active material.
The Zn2+ storage mechanism of VHCF was investigated in depth by XRD characterization and XPS analysis. Figure 3a shows the XRD curves of VHCF at different charging/discharging stages and Figure 3b depicts the (200) interplanar spacing variations calculated by the Bragg formula. It can be observed that the (200) diffraction peak shifts slightly to a lower angle with interplanar spacing increasing from 5.07 Å to 5.08 Å after first being charged to 2.0 V along with the de-intercalation of K+ ions. During the subsequent discharge and charge, the interplanar spacing shrinks from 5.08 Å to 5.03 Å and then returns to the value after the first charge, indicating the reversible structure change of the VHCF electrode. It can be found that the (200) interplanar spacing expansions and contractions are very small (~1%), with almost zero strain [30]. XPS was employed to further investigate the valence change of VHCF after the electrochemical reactions. As shown in Figure S2, the intensity of the K XPS peak decreases after being charged to 2.0 V, indicating that part of the K+ ions (not all of them) are extracted from the VHCF framework during the initial charging process. When the electrode is discharged to 0.2 V, the intensity of the K XPS peak is almost unchanged. It reveals that K+ will not participate in the following electrochemical reactions. Figure 3c shows the Fe 2p XPS region. It can be found that Fe2+ in VHCF is oxidized to Fe3+ during the first charge and reduced to Fe2+ after the first discharge. In addition, V can be another redox-active site of VHCF. As shown in Figure 3d, the initial valence of V in the VHCF electrode is 4+. When charged to 2.0 V, V4+ is partially oxidized to V5+, and the average valence of V becomes 4.72+. A new XPS peak appears at 516.7 eV after discharging to 0.2 V, which can be assigned to V3+, indicating that V is reduced with the intercalation of zinc ions. As shown in Figure S3, three charge/discharge plateaus can be observed in a typical charge/discharge profile. Plateau 1 between 1.2 V and 2.0 V is attributed to the redox of Fe ions. Plateau 2 from 0.6 V to 1.2 V and plateau 3 from 0.2 V to 0.6 V are attributed to the redox of V ions. The corresponding capacities from the Fe redox center (70 mA h g−1) and V redox center (160 mA h g−1) are consistent with the ratio of Fe and V in the chemical formula of VHCF. These results demonstrate that there are two electrochemical redox-active centers in the VHCF electrode (Fe2+ ⇋ Fe3+ and V5+ ⇋ V4+ ⇋ V3+) [31,32]. This enables the Zn//VHCF battery to achieve a three-electron-involved redox reaction, providing an ultra-high reversible capacity.
Although the VHCF cathode exhibits the advantages of high capacity and high voltage, its very bad cycling performance means that it is not competitive with other materials. The dissolution of active species in a dilute aqueous electrolyte is a common problem for PBA cathodes. To enhance the cycling performance of VHCF, a high-concentration hybrid electrolyte (3 M Zn(CF3SO3)2 + 10 M LiTFSI) was used. Figure 4a compares the dissolution of a VHCF electrode after immersion in two different electrolytes at room temperature for 2 weeks. It can be found that the 3 M Zn(CF3SO3)2 electrolyte turns yellow, while the 3 M Zn(CF3SO3)2 + 10 M LiTFSI electrolyte is still almost colorless and transparent. This demonstrates that the high-concentration electrolyte can prevent the dissolution of VHCF. As plotted in Figure 4b, the discharge capacities and the coulombic efficiencies of the first three cycles are 126.1 mA h g−1, 131.9 mA h g−1, 140.2 mA h g−1, and 99.3%, 96.3%, 97.0%, respectively. As shown in Figure 4c, the specific capacity of the VHCF electrode reaches the highest value of 228.8 mA h g−1 in 45 cycles. A capacity of 214 mA h g−1 is maintained after 100 cycles (93.5%). Although the first discharge capacity of VHCF in this electrolyte is lower, it can still reach almost the same capacity as that in 3 M Zn(CF3SO3)2 after an activation process. Figure S4 shows the charge/discharge profiles of VHCF in the 3 M Zn(CF3SO3)2 + 10 M LiTFSI electrolyte at 0.2 A g−1 after 20, 40, 60, 80, and 100 cycles, respectively. It can be found that the increase in capacity before the 45th cycle is mainly due to the elongation of the plateaus between 0.2 V and 1.2 V. It indicates that the redox reactions of V ions in VHCF are gradually activated during the initial cycling stage. However, the rate capability of the cell is slightly reduced, mainly caused by the lower ion conductivity of the high-concentration electrolyte (Figure 4d). The recorded capacities are 171.0, 194.7, 191.5, 166.1, 143.2, and 100.9 mA h g−1 at 0.2, 0.5, 1.0, 2.0, 3.0, and 5.0 A g−1, respectively. Nevertheless, its energy/power densities at 5.0 A g−1 (81.0 Wh Kg−1/4013.9 W Kg−1) are still higher than that of the traditional PBA cathodes in AZIBs [33,34,35]. Moreover, it shows a high capacity retention of 92.9% with an average efficiency of ~100% after 1000 cycles (Figure 4e). As shown in Figure S5, the surfaces of the VHCF electrodes are covered with a film after 1000 cycles, which may be a solid electrolyte interface (SEI) film. It can be found that the film on the VHCF electrode surface in 3 M Zn(CF3SO3)2 +10 M LiTFSI after 1000 cycles is denser and more uniform, which can more efficiently inhibit the dissolution of active material. The above results suggest that the Zn//VHCF batteries using the high-concentration hybrid electrolyte have the advantages of high capacity, high voltage, and excellent long-cycling stability, and are expected to become the next generation of safe and cheap large-scale energy storage batteries.

4. Conclusions

In summary, VHCF powders can be easily prepared and are suitable for AZIBs as the cathode material. The VHCF cathode exhibits an impressive capacity of 230.3 mA h g−1 and a high average voltage of ~1.2 V. Experimental investigations demonstrate that the high capacity comes from the two electrochemical redox-active centers in VHCF (Fe2+ ⇋ Fe3+ and V5+ ⇋ V4+ ⇋ V3+), and the intercalation/de-intercalation of Zn2+ can occur without VHCF phase transition at almost zero lattice strain. The fast capacity decay in the 3 M Zn(CF3SO3)2 electrolyte is due to the loss of VHCF. The cycling performance of the Zn//VHCF cell can be largely improved by using a high-concentration hybrid electrolyte that can inhibit the dissolution of VHCF. The improved battery exhibits a high capacity of 235.8 mA h g−1 and a remarkable capacity retention of 92.9% after 1000 cycles. These findings not only contribute to understanding the electrochemical behaviors of PBA cathodes in AZIBs, but also have important implications for the development of more high-performance cathode materials for cost-efficient and safe energy storage systems. However, the detailed electrochemical mechanisms of VHCF, especially the structure evolution and the role of K+ ions during the charge/discharge process require further studies. These issues will become a focus of our future research.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano12234268/s1, Figure S1: The cycling performance of the Zn//VHCF battery in 3 M Zn(CF3SO3)2 electrolyte at 0.2 A g−1 with a voltage window between 1.2 V and 2.0 V; Figure S2: K 1s XPS regions for the pristine, charged and discharged VHCF electrodes; Figure S3: The charge/discharge profiles of the VHCF//Zn battery in 3 M Zn(CF3SO3)2 electrolyte after 15 cycles at 0.2 A g−1; Figure S4: The charge/discharge profiles of VHCF in 3 M Zn(CF3SO3)2 + 10 M LiTFSI electrolyte at 0.2 A g−1 after 20, 40, 60, 80 and 100 cycles, respectively; Figure S5: The FESEM images of (a) the pristine VHCF electrodes, (b) the VHCF electrode in 3 M Zn(CF3SO3)2 after 1000 cycles, and (c) the VHCF electrode in 3 M Zn(CF3SO3)2 +10 M LiTFSI after 1000 cycles.

Author Contributions

Conceptualization, S.Z., Q.P. and Y.T.; methodology, S.Z. and Q.P.; software, S.Z.; validation, Q.P., Y.T. and X.L.; formal analysis, S.Z.; investigation, W.H.; resources, X.L. and Y.F.; data curation, S.Z. and Y.A.; writing—original draft preparation, S.Z.; writing—review and editing, Q.P., Y.T., M.X., X.L. and S.Z.; visualization, S.Z.; supervision, X.L. and Q.P.; project administration, X.L.; funding acquisition, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 21905036), the Fundamental Research Funds for the Central Universities (Grant No. 017192610, 3132022195), and the China Postdoctoral Science Foundation (Grant No. 2019M651198).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant No. 21905036), the Fundamental Research Funds for the Central Universities (Grant No. 017192610, 3132022195), and the China Postdoctoral Science Foundation (Grant No. 2019M651198).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, K.; Liu, Y.Y.; Lin, D.C.; Pei, A.; Cui, Y. Materials for Lithium-Ion Battery Safety. Sci. Adv. 2018, 4, eaas9820. [Google Scholar] [CrossRef] [Green Version]
  2. Liu, B.H.; Jia, Y.K.; Yuan, C.H.; Wang, L.B.; Gao, X.; Yin, S.; Xu, J. Safety Issues and Mechanisms of Lithium-Ion Battery Cell Upon Mechanical Abusive Loading: A Review. Energy Storage Mater. 2020, 24, 85–112. [Google Scholar] [CrossRef]
  3. Yang, C.; Xin, S.; Mai, L.Q.; You, Y. Materials Design for High-Safety Sodium-Ion Battery. Adv. Energy Mater. 2021, 11, 2000974. [Google Scholar] [CrossRef]
  4. Pang, Q.; Yu, X.Y.; Zhang, S.J.; He, W.; Yang, S.Y.; Fu, Y.; Tian, Y.; Xing, M.M.; Luo, X.X. High-Capacity and Long-Lifespan Aqueous LiV3O8/Zn Battery Using Zn/Li Hybrid Electrolyte. Nanomaterials. 2021, 11, 1429. [Google Scholar] [CrossRef]
  5. Yu, P.; Zeng, Y.X.; Zhang, H.Z.; Yu, M.H.; Tong, Y.X.; Lu, X.H. Flexible Zn-Ion Batteries: Recent Progresses and Challenges. Small 2019, 15, 1804760. [Google Scholar] [CrossRef]
  6. Harting, K.; Kunz, U.; Turek, T. Zinc-Air Batteries: Prospects and Challenges for Future Improvement. Z. Phys. Chemie Int. J. Res. Phys. Chem. Chem. Phys. 2012, 226, 151–166. [Google Scholar] [CrossRef]
  7. Shi, Y.C.; Chen, Y.; Shi, L.; Wang, K.; Wang, B.; Li, L.; Ma, Y.M.; Li, Y.H.; Sun, Z.H.; Ali, W.; et al. An Overview and Future Perspectives of Rechargeable Zinc Batteries. Small 2020, 16, 2000730. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, F.; Borodin, O.; Gao, T.; Fan, X.L.; Sun, W.; Han, F.D.; Faraone, A.; Dura, J.A.; Xu, K.; Wang, C.S. Highly Reversible Zinc Metal Anode for Aqueous Batteries. Nat. Mater. 2018, 17, 543–549. [Google Scholar] [CrossRef]
  9. Wu, T.-H.; Zhang, Y.; Althouse, Z.D.; Liu, N. Nanoscale Design of Zinc Anodes for High-Energy Aqueous Rechargeable Batteries. Mater. Today Nano 2019, 6, 100032. [Google Scholar] [CrossRef]
  10. Tan, Y.; An, F.Q.; Liu, Y.C.; Li, S.W.; He, P.G.; Zhang, N.; Li, P.; Qu, X.H. Reaction Kinetics in Rechargeable Zinc-Ion Batteries. J. Power Sources 2021, 492, 229655. [Google Scholar] [CrossRef]
  11. Zuo, Y.; Wang, K.; Pei, P.; Wei, M.; Liu, X.; Xiao, Y.; Zhang, P. Zinc Dendrite Growth and Inhibition Strategies. Mater. Today Energy 2021, 20, 100692. [Google Scholar] [CrossRef]
  12. Wu, J.; Kuang, Q.; Zhang, K.; Feng, J.J.; Huang, C.M.; Li, J.J.; Fan, Q.H.; Dong, Y.Z.; Zhao, Y.M. Spinel Zn3V3O8: A High-Capacity Zinc Supplied Cathode for Aqueous Zn-Ion Batteries. Energy Storage Mater. 2021, 41, 297–309. [Google Scholar] [CrossRef]
  13. Yang, B.; Cao, X.W.; Wang, S.H.; Wang, N.; Sun, C.L. Manganese Oxides Hierarchical Microspheres as Cathode Material for High-Performance Aqueous Zinc-Ion Batteries. Electrochim. Acta 2021, 385, 138447. [Google Scholar] [CrossRef]
  14. Xiao, X.C.; Ahn, D.; Liu, Z.Y.; Kim, J.-H.; Lu, P. Atomic Layer Coating to Mitigate Capacity Fading Associated with Manganese Dissolution in Lithium Ion Batteries. Electrochem. Commun. 2013, 32, 31–34. [Google Scholar] [CrossRef]
  15. Abbas, S.; Mehboob, S.; Shin, H.-J.; Han, O.H.; Ha, H.Y. Highly Functionalized Nanoporous Thin Carbon Paper Electrodes for High Energy Density of Zero-Gap Vanadium Redox Flow Battery. Chem. Eng. J. 2019, 378, 122190. [Google Scholar] [CrossRef]
  16. Chen, J.S.; Wei, L.; Mahmood, A.; Pei, Z.X.; Zhou, Z.; Chen, X.C.; Chen, Y. Prussian Blue, Its Analogues and Their Derived Materials for Electrochemical Energy Storage and Conversion. Energy Storage Mater. 2020, 25, 585–612. [Google Scholar] [CrossRef]
  17. Li, W.-J.; Han, C.; Cheng, G.; Chou, S.-L.; Liu, H.-K.; Dou, S.-X. Chemical Properties, Structural Properties, and Energy Storage Applications of Prussian Blue Analogues. Small 2019, 15, 1900470. [Google Scholar] [CrossRef]
  18. Qian, J.F.; Wu, C.; Cao, Y.L.; Ma, Z.F.; Huang, Y.H.; Ai, X.P.; Yang, H.X. Prussian Blue Cathode Materials for Sodium-Ion Batteries and Other Ion Batteries. Adv. Energy Mater. 2018, 8, 1702619. [Google Scholar] [CrossRef]
  19. Wang, B.Q.; Han, Y.; Wang, X.; Bahlawane, N.; Pan, H.G.; Yan, M.; Jiang, Y.Z. Prussian Blue Analogs for Rechargeable Batteries. iScience 2018, 3, 110–133. [Google Scholar] [CrossRef] [Green Version]
  20. Bie, X.F.; Kubota, K.; Hosaka, T.; Chihara, K.; Komaba, S. Synthesis and Electrochemical Properties of Na-Rich Prussian Blue Analogues Containing Mn, Fe, Co, and Fe for Na-Ion Batteries. J. Power Sources 2018, 378, 322–330. [Google Scholar] [CrossRef]
  21. Ge, P.; Li, S.J.; Shuai, H.L.; Xu, W.; Tian, Y.; Yang, L.; Zou, G.Q.; Hou, H.S.; Ji, X.B. Ultrafast Sodium Full Batteries Derived from X-Fe (X = Co, Ni, Mn) Prussian Blue Analogs. Adv. Mater. 2019, 31, 1806092. [Google Scholar] [CrossRef]
  22. Zhang, L.Y.; Chen, L.; Zhou, X.F.; Liu, Z.P. Towards High-Voltage Aqueous Metal-Ion Batteries Beyond 1.5 V: The Zinc/Zinc Hexacyanoferrate System. Adv. Energy Mater. 2015, 5, 1400930. [Google Scholar] [CrossRef]
  23. Gorlin, M.; Ojwang, D.O.; Lee, M.-T.; Renman, V.; Tai, C.-W.; Valvo, M. Aging and Charge Compensation Effects of the Rechargeable Aqueous Zinc/Copper Hexacyanoferrate Battery Elucidated Using in Situ X-Ray Techniques. ACS Appl. Mater. Interfaces 2021, 13, 59962–59974. [Google Scholar] [CrossRef] [PubMed]
  24. Xue, Y.T.; Chen, Y.; Shen, X.P.; Zhong, A.; Ji, Z.Y.; Cheng, J.; Kong, L.R.; Yuan, A.H. Decoration of Nickel Hexacyanoferrate Nanocubes onto Reduced Graphene Oxide Sheets as High-Performance Cathode Material for Rechargeable Aqueous Zinc-Ion Batteries. J. Colloid Interface Sci. 2022, 609, 297–306. [Google Scholar] [CrossRef] [PubMed]
  25. Ma, L.T.; Chen, S.M.; Long, C.B.; Li, X.L.; Zhao, Y.W.; Liu, Z.X.; Huang, Z.D.; Dong, B.B.; Zapien, J.A.; Zhi, C.Y. Achieving High-Voltage and High-Capacity Aqueous Rechargeable Zinc Ion Battery by Incorporating Two-Species Redox Reaction. Adv. Energy Mater. 2019, 9, 1902446. [Google Scholar] [CrossRef]
  26. Peng, X.C.; Guo, H.C.; Ren, W.H.; Su, Z.; Zhao, C. Vanadium Hexacyanoferrate as High-Capacity Cathode for Fast Proton Storage. Chem. Commun. 2020, 56, 11803–11806. [Google Scholar] [CrossRef]
  27. Xiong, P.X.; Zeng, G.J.; Zeng, L.X.; Wei, M.D. Prussian Blue Analogues Mn[Fe(CN)6]0.6667·nH2O Cubes as an Anode Material for Lithium-Ion Batteries. Dalton Trans. 2015, 44, 16746–16751. [Google Scholar] [CrossRef]
  28. Xu, C.J.; Li, B.H.; Du, H.D.; Kang, F.Y. Energetic Zinc Ion Chemistry: The Rechargeable Zinc Ion Battery. Angew. Chem. Int. Edit. 2012, 51, 933–935. [Google Scholar] [CrossRef]
  29. Yan, M.Y.; He, P.; Chen, Y.; Wang, S.Y.; Wei, Q.L.; Zhao, K.N.; Xu, X.; An, Q.Y.; Shuang, Y.; Shao, Y.Y.; et al. Water-Lubricated Intercalation in V2O5·nH2O for High-Capacity and High-Rate Aqueous Rechargeable Zinc Batteries. Adv. Mater. 2018, 30, 1703725. [Google Scholar] [CrossRef]
  30. Jiang, P.; Lei, Z.Y.; Chen, L.; Shao, X.C.; Liang, X.M.; Zhang, J.; Wang, Y.C.; Zhang, J.J.; Liu, Z.P.; Feng, J.W. Polyethylene Glycol-Na+ Interface of Vanadium Hexacyanoferrate Cathode for Highly Stable Rechargeable Aqueous Sodium-Ion Battery. ACS Appl. Mater. Interfaces 2019, 11, 28762–28768. [Google Scholar] [CrossRef]
  31. Wan, F.; Niu, Z.Q. Design Strategies for Vanadium-Based Aqueous Zinc-Ion Batteries. Angew. Chem. Int. Edit. 2019, 58, 16358–16367. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, N.; Dong, Y.; Jia, M.; Bian, X.; Wang, Y.Y.; Qiu, M.D.; Xu, J.Z.; Liu, Y.C.; Jiao, L.F.; Cheng, F.Y. Rechargeable Aqueous Zn-V2O5 Battery with High Energy Density and Long Cycle Life. ACS Energy Lett. 2018, 3, 1366–1372. [Google Scholar] [CrossRef]
  33. Trocoli, R.; La Mantia, F. An Aqueous Zinc-Ion Battery Based on Copper Hexacyanoferrate. ChemSusChem 2015, 8, 481–485. [Google Scholar] [CrossRef]
  34. Jia, Z.J.; Wang, B.G.; Wang, Y. Copper Hexacyanoferrate with a Well-Defined Open Framework as a Positive Electrode for Aqueous Zinc Ion Batteries. Mater. Chem. Phys. 2015, 149, 601–606. [Google Scholar] [CrossRef]
  35. Renman, V.; Ojwang, D.O.; Valvo, M.; Gómez, C.P.; Gustafsson, T.; Svensson, G. Structural-Electrochemical Relations in the Aqueous Copper Hexacyanoferrate-Zinc System Examined by Synchrotron X-Ray Diffraction. J. Power Sources 2017, 369, 146–153. [Google Scholar] [CrossRef]
Figure 1. (a) XRD pattern, (b) SEM image, (c) TEM image, (d) TG curves, and (e) EDS images of VHCF.
Figure 1. (a) XRD pattern, (b) SEM image, (c) TEM image, (d) TG curves, and (e) EDS images of VHCF.
Nanomaterials 12 04268 g001
Figure 2. Electrochemical performance of VHCF in the 3 M Zn(CF3SO3)2 electrolyte; (a) The initial three charge/discharge profiles; (b) Cycling performance; (c) Rate performance; (d) Long-cycling stability.
Figure 2. Electrochemical performance of VHCF in the 3 M Zn(CF3SO3)2 electrolyte; (a) The initial three charge/discharge profiles; (b) Cycling performance; (c) Rate performance; (d) Long-cycling stability.
Nanomaterials 12 04268 g002
Figure 3. (a) XRD curves and (b) (200) interplanar spacing variations of VHCF at different charge/discharge stages of the first cycle; (c) Fe 2p and (d) V 2p XPS regions for the pristine, charged, and discharged VHCF electrodes.
Figure 3. (a) XRD curves and (b) (200) interplanar spacing variations of VHCF at different charge/discharge stages of the first cycle; (c) Fe 2p and (d) V 2p XPS regions for the pristine, charged, and discharged VHCF electrodes.
Nanomaterials 12 04268 g003
Figure 4. (a) Digital photo of the VHCF electrodes after immersion in two different electrolytes for 2 weeks at room temperature; (b) The initial three charge/discharge profiles; (c) cycling performance; (d) rate performance; and (e) long-cycling stability of VHCF in the 3 M Zn(CF3SO3)2 + 10 M LiTFSI electrolyte.
Figure 4. (a) Digital photo of the VHCF electrodes after immersion in two different electrolytes for 2 weeks at room temperature; (b) The initial three charge/discharge profiles; (c) cycling performance; (d) rate performance; and (e) long-cycling stability of VHCF in the 3 M Zn(CF3SO3)2 + 10 M LiTFSI electrolyte.
Nanomaterials 12 04268 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, S.; Pang, Q.; Ai, Y.; He, W.; Fu, Y.; Xing, M.; Tian, Y.; Luo, X. Vanadium Hexacyanoferrate as a High-Capacity and High-Voltage Cathode for Aqueous Rechargeable Zinc Ion Batteries. Nanomaterials 2022, 12, 4268. https://doi.org/10.3390/nano12234268

AMA Style

Zhang S, Pang Q, Ai Y, He W, Fu Y, Xing M, Tian Y, Luo X. Vanadium Hexacyanoferrate as a High-Capacity and High-Voltage Cathode for Aqueous Rechargeable Zinc Ion Batteries. Nanomaterials. 2022; 12(23):4268. https://doi.org/10.3390/nano12234268

Chicago/Turabian Style

Zhang, Shijing, Qiang Pang, Yuqing Ai, Wei He, Yao Fu, Mingming Xing, Ying Tian, and Xixian Luo. 2022. "Vanadium Hexacyanoferrate as a High-Capacity and High-Voltage Cathode for Aqueous Rechargeable Zinc Ion Batteries" Nanomaterials 12, no. 23: 4268. https://doi.org/10.3390/nano12234268

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop